Next Article in Journal
Multi-Parameter Experimental Investigation on the Characteristics of Acidizing Effectiveness in High-Temperature Carbonate Formation
Previous Article in Journal
An Eight-Coil Wireless Power Transfer Method for Improving the Coupling Tolerance Based on Uniform Magnetic Field
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Fe2NiSe4 Nanowires Array for Highly Efficient Electrochemical H2S Splitting and Simultaneous Energy-Saving H2 Production

1
Research Institute of Natural Gas Technology, PetroChina Southwest Oil & Gas Field Company, Chengdu 610213, China
2
National R&D Center for High Sulfur Gas Exploitation, Chengdu 610000, China
3
Key Laboratory of Natural Gas Quality Control and Energy Measurement for State Market Regulation, Chengdu 610213, China
4
PetroChina Southwest Oil & Gas Field Company, Chengdu 610051, China
5
School of New Energy and Materials, Southwest Petroleum University, Chengdu 610500, China
*
Author to whom correspondence should be addressed.
Processes 2024, 12(10), 2111; https://doi.org/10.3390/pr12102111 (registering DOI)
Submission received: 15 April 2024 / Revised: 15 May 2024 / Accepted: 21 May 2024 / Published: 27 September 2024
(This article belongs to the Section Energy Systems)

Abstract

:
The electrochemical removal of abundant and toxic H2S from highly sour reservoirs has emerged as a promising method for hydrogen production and desulfurization. Nevertheless, the ineffectiveness and instability of current electrocatalysts have impeded further utilization of H2S. In this communication, we introduce a robust array of Fe2NiSe4 nanowires synthesized in situ on a FeNi3 foam (Fe2NiSe4/FeNi3) via hydrothermal treatment. This array acts as an active electrocatalyst for ambient H2S splitting. It offers numerous exposed active sites and a rapid electron transport channel, significantly enhancing charge transport rates. As an electrode material, Fe2NiSe4/FeNi3 displays remarkable electrocatalytic efficiency for both sulfide oxidation and hydrogen evolution reactions. This bifunctional electrode achieves efficient electrochemical H2S splitting at a low potential of 440 mV to reach a current density of 100 mA∙cm−2, with a faradaic efficiency for hydrogen production of approximately 98%. These findings highlight its significant potential for desulfurization and energy-efficient hydrogen generation.

1. Introduction

With the continuous advancement of the energy revolution, energy consumption is evolving toward electrification, cleanliness, low carbon, and intelligence. At present, it is a critical period to accelerate energy transformation and promote “carbon peak” and “carbon neutrality”. Natural gas, as the cleanest and lowest-carbon fossil fuel, plays a crucial role in transitioning from high-carbon to low-carbon energy sources and has been extensively utilized for reducing carbon emissions in recent decades [1]. However, the focus on natural gas has shifted from conventional gas reservoirs to sour ones, in which the H2S reached 10 million tons per year worldwide in 2019 [2]. The substantial amount of H2S, with its strong toxicity, corrosivity, and malodor, has traditionally been deemed useless or even hazardous, requiring pretreatment before further processing. Efficient desulfurization through sulfide oxidation reactions is crucial for environmentally friendly development, as well as for enhancing human health and safety [3]. Traditional desulfurization methods, like chemical adsorption, biological oxidation, or catalytic conversion, demand substantial energy and chemical use, making them energy-intensive. The Claus process, the standard industrial method operating at high temperatures (900 °C), partially oxidizes H2S to produce sulfur and water vapor. Yet, this process only recycles sulfur, with the potential energy of hydrogen being lost as water vapor, leading to the underutilization of H2S resources.
Hydrogen energy, as a clean and sustainable new energy source, has a wide range of application scenarios. For example, it is utilized in distributed power generation or cogeneration to provide heat and power buildings [4,5], as hydrogen fuel for fuel cell vehicles [6], as reducing agents for industrial metallurgy, and as raw materials for the synthesis of methanol, ammonia, and other chemical products [7]. Electrochemical water splitting driven by sustainable energy has been investigated as a prospective alternative way to produce green hydrogen, especially with the rapid growth of clean energy power capacity [8,9]. However, commercial water electrolyzers typically operate at a much larger voltage of 1.8 to 2.0 V, especially the oxygen evolution reaction (OER), with high activation energy barriers for O-O bond formation, and sustain huge overall electricity consumption for the multiple proton and electron transfer steps [10]. Therefore, replacing the oxygen evolution reaction with a thermodynamically more readily oxidation reaction to reduce overall energy consumption offers a viable strategy for an energy-saving approach to facilitate H2 production. Tang et al. developed a Ni2P nanoarray as a high-performance non-noble-metal electrocatalyst for anodic hydrazine oxidation reaction and cathodic for low-energy hydrogen production [11]. Moreover, the anode oxidation reactions have the extra advantage of obtaining valuable chemicals or dealing with environmental pollution.
Electrochemical conversion of H2S into hydrogen and sulfur products is an economically and environmentally favorable method, addressing significant industrial waste gas issues. Electrocatalytic H2S offers a mild and efficient means of hydrogen production and sulfur recycling. From the thermodynamic point of view, H2S splitting requires less energy compared to water splitting (33 kJ mol−1 for H2S, 237 kJ mol−1 for H2O), suggesting its feasibility. Nevertheless, its effectiveness has been limited by the inefficient activity and inadequate stability of electrocatalysts in sulfur-rich environments [12]. To solve the problem, an indirect electrochemical–chemical loop strategy has been developed using the intermediate redox media to oxidize sulfide anions. Although this method has achieved modest success in preventing direct contact of the electrocatalysts with sulfur-containing solutions, it faces challenges such as the need for complex redox reagents (Fe2+/Fe3+, VO2+/VO2+, I/I3) and poor long-term stability in acidic conditions. Alternatively, direct electrocatalytic H2S splitting methods have also been developed. For example, a series of metal sulfide catalysts have shown sulfur resistance under harsh electrolyte conditions, making them potential electrocatalysts [13,14,15,16,17,18,19]. However, its potential problems, such as poor electrical conductivity, restrict its further development [20,21]. The metal selenides also demonstrate excellent stability, great conductivity, and high activity in sulfur-rich conditions. Duan et al. reported an anti-sulfuretted NiSe/NF array catalyst for sulfur oxidation reaction with low-energy H2 production, exhibiting 500 h outstanding stability. However, further improvement in its catalytic activity remains necessary [22].
Metal selenides and nanostructure materials have received extensive attention due to their great electrochemical properties [20,23,24]. Motivated by these advancements, we developed a Fe2NiSe4 nanowires array grown in situ on conductive FeNi3 foam substrates (Fe2NiSe4/FeNi3). This configuration exposes numerous active sites and facilitates rapid electron transport, enabling high-efficiency electrocatalytic decomposition of H2S coupled with hydrogen production. Remarkably, the Fe2NiSe4/FeNi3 serves as an attractive bifunctional catalyst for hydrogen generation and sulfide oxidation reaction (SOR), outperforming previously reported anti-sulfur NiSe/NF catalysts. It achieves an impressive performance in alkaline environments, requiring a low potential of 440 mV to achieve 100 mA∙cm−2. Additionally, it demonstrates robust durability for more than 800 min and approximately 98% faradaic efficiency toward H2 production. Furthermore, the product of anodic sulfide oxidation follows the chain-growing mechanism of S2−/HS to polysulfides in alkaline solutions by combining UV–vis analysis. Our research provides guidance for designing efficient electrocatalysts of metal selenide catalysts based on highly efficient electrochemical H2S splitting and simultaneous energy-saving H2 production.

2. Materials and Methods

2.1. Chemicals

The commercial FeNi3 foam was purchased from Kunshan Long Sheng Bao electronic material form Su Zhou city. Reagents used in the experiment were purchased from Shanghai Aladdin Industrial Inc. (China) with the following purities: NaBH4 (97%), Se powder (99.9%), HCl (36–38%), ethanol (75%), Na2S·9H2O (98%), and NaOH (98%). All reagents were of analytical pure grade and were without any treatment before use. All experimental water was used ultrapure water through a Millipore system.

2.2. Synthesis of Fe2NiSe4/FeNi3

An Fe2NiSe4 nanowires array was synthesized on FeNi3 foam through a one-step hydrothermal method. The synthesis process is shown in Figure S1. First, prepare 2 × 3 cm2 of FeNi3 foam and wash it several times with 3 M HCl, ethanol, and deionized water to ensure that the surface of FeNi3 foam was adequately cleaned before use. Firstly, to prepare the NaHSe solution, 0.118 g Se powder was added to 3 mL ultrapure water along with 0.130 g NaBH4. The mixture was stirred continuously until it became clear. Next, the prepared NaHSe solution was added to a 30 mL ethanol solution. The resulting mixture was then transferred into a 50 mL Teflon-lined stainless steel autoclave with previously treated FeNi3 inside. The autoclave was sealed tightly and placed in an oven at 180 °C for 12 h. After the reaction, we allowed the autoclave to slowly cool to room temperature. The post-reaction sample was rinsed with ultrapure water and ethanol and, finally, placed in a vacuum drying oven at 60 °C for 4 h. As a control material, NiSe was synthesized according to the previous literature [22].

2.3. Characterization

X-ray diffraction (XRD) data were acquired on a Philips X′Pert diffractometer with Cu Kα radiation (λ = 1.5418 Å). Scanning electron microscopy (SEM) measurements were carried out on a ZEISS Sigma 300 scanning electron microscope at an accelerating voltage of 20 kV. Transmission electron microscopy (TEM) measurements were performed on an FEI TECNAI G20 electron microscope (FEI, American) with an accelerating voltage of 200 kV. Energy dispersive spectroscopy (EDS) mapping was conducted by FEI TF20. X-ray photoelectron spectroscopy (XPS) measurements were performed on a Thermo Fisher Scientific K-Alpha X-ray photoelectron spectrometer using Mg as the exciting source. UV-vis absorption spectra were recorded with a Shimadzu UV-2600 spectrophotometer (Shimadzu, Japan)

2.4. Electrochemical Measurement

All electrochemical measurements were performed with a CHI 660D electrochemical analyzer (CH660 Instruments, Inc., Shanghai, China) in a standard three-electrode system using Fe2NiSe4/FeNi3 as the working electrode, a graphite rod as the counter electrode, and a Hg/HgO electrode as the reference electrode. All electrochemical tests were carried out in a 50 mL H-cell. The electrolyte solution of 1.0 M NaOH with 1.0 M Na2S, as a simulated hydrogen sulfide absorption solution, served as an electrolyte for the half-cell SOR study, while the electrolyte solution of 1.0 M NaOH served for the half-cell hydrogen evolution reaction and half-cell oxygen evolution reaction. All potentials measured were calibrated to RHE using the following equation: E (RHE) = E (Hg/HgO) + 0.098 V + 0.0591 × pH. Polarization curves were obtained using linear sweep voltammetry (LSV) with a scan rate of 5 mV s−1, and no activation was used before recording the polarization curves. Electrochemical impedance spectra (EIS) were carried out from 1 Hz to 1000 KHz at the open circuit voltage. Electrochemical active surface area (ECSA) was calculated by capacitance (Cdl) measurements. The long-term durability was performed using chronoamperometry measurements.

3. Results and Discussion

3.1. Characterization of Fe2NiSe4/FeNi3

The catalytic material of Fe2NiSe4/FeNi3 was successfully prepared by the hydrothermal synthesis method. Figure 1a displays the XRD patterns of Fe2NiSe4/FeNi3 and FeNi3. Three strong diffraction peaks at 44.5°, 51.8°, and 76.1° belong to the (111), (200), and (220) crystal faces of FeNi3, respectively (JCPDS no. 38-0419) [25]. After selenization of the FeNi3 foam, a new series of diffraction peaks at 28.9°, 33.1°, 59.6°, and 70.6° corresponded to (220), (−112), (116), and (404) crystal faces of Fe2NiSe4 solid solution materials, respectively (JCPDS no. 65-2338) [26]. The surface morphologies of Fe2NiSe4/FeNi3 were characterized by scanning electron microscope and transmission electron microscopy images. From Figure 1b–d and Figure S2, it is evident that the Fe2NiSe4/FeNi3 exhibits the shape of the nanowires array. Figure 1e shows the high-resolution transmission electron microscopy (HRTEM) image, which displays well-defined lattice fringes with interplanar distances of 2.6 Å and 2.7 Å, respectively. The angle between the two crystal faces is 73°. This corresponds to the Fe2NiSe4 (213) and (−303) crystal planes, respectively. At the same time, the EDS mapping images of Fe2NiSe4 nanowire further reveal the uniform distribution of Fe, Ni, and Se.
The valence state of Fe2NiSe4/FeNi3 was determined using X-ray photoelectron spectroscopy. The survey spectrum shown in Figure S3 confirms the characteristic peaks for Ni, Fe, and Se, and the C and O signals may be attributed to air contamination. Figure 2a displays the high-resolution XPS spectrum of the Ni 2p region. Two spin-orbit doublets characteristic of peaks around 853 eV and 871 eV are attributed to Ni 2p3/2 and Ni 2p1/2, respectively. The binding energies of 876.9 eV and 858.9 eV are ascribed to two shakeup satellites from superficial oxidation of Fe2NiSe4 owing to air contact. The other binding energies of 870.9 eV and 853.2 eV are attributed to Ni2+ [27]. Figure 2b shows the high-resolution Fe 2p spectrum. The peak at 710.2 eV is indexed to Fe 2p3/2, and the peak at 723.8 eV is distributed to Fe 2p1/2 [28,29]. As for the Se 3d spectrum in Figure 2c, the binding energies of 54.72 and 53.71 eV are assigned to Se 3d3/2 and Se 3d5/2, respectively, which are assigned to the Se bonded to Ni or Fe in the form of metal selenide [28,29]. The peak at 58.32 eV is associated with the surface oxidation state of Se species. All these observations confirm the successful formation of Fe2NiSe4 on FeNi3 foam.

3.2. Electrocatalytic Performance

The electrocatalytic performance of Fe2NiSe4/FeNi3 for electrochemical H2S splitting was examined in 1.0 M NaOH solution with 1.0 M Na2S, employing a three-electrode system: a Hg/HgO as the reference electrode, a graphite rod as the counter electrode, and a square Fe2NiSe4/FeNi3 with a side length of 0.5 * 0.5 cm as the work electrode. An H-type electrolytic cell with a separate cathode and anode was used for the electrocatalytic activity study. The image of the laboratory electrochemical cell is shown in Figure S4. Figure 3a illustrates the LSV polarization curves of Fe2NiSe4/FeNi3, NiSe/NF, and bare FeNi3 foam with a scan rate of 5 mv s−1 for SOR. It can be found that bare FeNi3 foam exhibits poor SOR activity. In sharp contrast, the Fe2NiSe4/FeNi3 catalyst demonstrates approximately twice the catalytic activity compared to NiSe/NF in the SOR process, achieving a current density of 100 mA cm−2 at a potential of 0.44 V vs. RHE. This potential is 50 mV lower than that of NiSe/NF and compares favorably to the behaviors of transition metal sulfide catalysts (as shown in Table S1). These findings suggest that the Fe2NiSe4/FeNi3 electrode exhibits high efficiency in catalyzing the SOR, which is well verified under the comparison of different voltages (0.4 V, 0.5 V, and 0.6 V vs. RHE) in Figure 3b. At the same potential, the Fe2NiSe4/FeNi3 electrode demonstrates a higher current density, indicating a faster rate of sulfide oxidation reaction. Subsequently, the electrocatalytic performances of HER for Fe2NiSe4/FeNi3, NiSe/NF, Pt plate, and bare FeNi3 foam were also evaluated. As illustrated in Figure 3c, Fe2NiSe4/FeNi3 achieves a low overpotential of 0.14 V to achieve a current density of 20 mA cm−2, the best performance among several catalysts, superior to NiSe/NF and commercial Pt plate electrodes. Strikingly, Fe2NiSe4/FeNi3 provides a current density of 100 mA cm−2 at an overpotential of 0.31 V. The above electrochemical performance suggests that Fe2NiSe4/FeNi3 is an active catalyst toward both SOR and HER. Owing to the outstanding activity toward both HER and SOR, we used Fe2NiSe4/FeNi3 as both anode and cathode for hydrogen sulfide and water electrolysis in an H-cell, respectively. As shown in Figure 3d, it can be found that when the current density reaches 100 mA cm−2, the electrolytic splitting of H2S saves 1.1 V compared to water splitting.
To understand the possible origins of the superior performance of the Fe2NiSe4/FeNi3, the effective surface areas (ECSAs) of Fe2NiSe4/FeNi, NiSe/NF, and FeNi3 foam were estimated by measuring the capacitances of the double layer at the solid–liquid interface of these electrodes. ECSA can be estimated from the electrochemical bilayer capacitance (Cdl). The cyclic voltammetric curves for various catalysts with scan rates of 20, 40, 80, 120, 160, and 200 mV s−1 in the range of 0.27 to 0.37 V vs. RHE with non-faraday interval were tested. As shown in Figure S5, the capacitances of Fe2NiSe4/FeNi3, NiSe/NF, and FeNi3 foam are evaluated at 37.96 mF cm−2, 28.41 mF cm−2, and 3.15 mF cm−2, respectively, implying that Fe2NiSe4/FeNi3 exhibits enlarged electrochemically active surface area with the possibilities of exposing more active sites. Moreover, the impedance characteristics of different materials were investigated. A smaller semicircle in the high-frequency region reflects lower charge transfer resistance (Rct). As depicted in Figure 3e, the Nyquist plots indicate that Fe2NiSe4/FeNi3 exhibits the smallest semicircle radius, demonstrating the lowest charge transfer impedance. Therefore, it has better charge transport capability and faster catalytic dynamics.
It is essential to evaluate the long-term stability of a catalyst electrode to assess its quality for practical applications. We performed chronoamperometric measurements at 0.4 V constant potential for 800 min of Fe2NiSe4/FeNi3 catalyst. Figure 3f illustrates that the current density of the Fe2NiSe4/FeNi3 electrode remained relatively stable at about 28 mA cm−2 in NaOH with Na2S electrolyte. To further assess the stability, we also compared the LSV curves before and after the long SOR, as shown in Figure S6a. We observed a slight degradation in electrocatalytic performance after prolonged reaction, indicating that the formation of polysulfides can affect the performance of the electrocatalyst during long-term operation. In contrast, the catalytic performance was restored by stirring the electrolyte at 100 rpm to enhance mass transfer, as shown in Figure S6b. These findings indicate that polysulfides can affect the long-term performance of electrocatalysts. When polysulfides are redistributed into the electrolyte, the catalyst resumes high performance. The morphology of the catalyst post-reaction was examined. As shown in Figure S7, the nanowires array structure maintained its integrity. However, previous studies have reported that the electrode material surface is easily passivated by electro-oxidized product of elemental sulfur and thereby stops the electrocatalytic reaction. We explored the contact angle of different materials with liquid sulfur to verify their sulfurophilic or sulfurophobic properties. The results show that the contact angles of Fe2NiSe4/FeNi3, NiSe/NF, and FeNi3 for liquid sulfur are 126°, 105°, and 68°, respectively (Figure S8). The Fe2NiSe4/FeNi3 shows a larger contact angle, proving that Fe2NiSe4/FeNi3 is a sulfurphobic material. So, the Fe2NiSe4/FeNi3 nanowires array retained steady activity within the process of SOR.

3.3. Products Tests

To investigate the actual yield and reaction process of electrocatalytic decomposition of H2S, the post-electrolytic practical product was studied. Firstly, we used the external standard method to calibrate the reaction cell to quantify and then carried out a galvanostatic test at a current density of 100 mA cm−2. The gas produced at the cathode was confirmed by gas chromatography to be H2. As the reaction proceeded, the H2 production rate was monitored in real time by gas chromatography every 10 min. The calculation result indicates that the H2 production rate is maintained at about 0.68 mL min−1 cm−2, and the faradaic efficiency of H2 production achieves 98% (Figure 4a,b). For the anode, during the progress of the reaction, the color of the electrolyte changed from clear to yellow. The products were studied in the electrolyte by a UV–vis spectrophotometry system. The peak at 293 nm can be reduced to Sn2− (Figure 4c). With the longer electrolysis time, the signal assigned to Sn2− is a more obvious strength, revealing that S2−/HS has been selectively converted to Sn2−. Furthermore, we acidified the anode electrolyte to obtain a yellow powdered product (Figure 4d). The product characterization by XRD identified sulfur. Therefore, electrochemical splitting of H2S not only decreases energy consumption for desulfuration but also produces valuable H2 and sulfur.

4. Conclusions

In summary, electrocatalytic decomposition of H2S into high-value H2 and sulfur is a promising approach for the utilization of H2S resources. We successfully prepared a Fe2NiSe4/FeNi3 electrocatalyst that was demonstrated to be a highly active bifunctional catalyst for energy-saving electrolytic H2 generation for H2S electrolysis. For the anodic SOR, the electrode provides an extremely low onset potential of 0.33 V vs. RHE, and the current density reaches 100 mA cm−2 at 0.44 V vs. RHE. At the cathode, we measured the H2 generation rate of about 0.68 mL min−1 cm−2 at a current density of 100 mA cm−2, and the faradaic efficiency of H2 was around 98%. This study demonstrates that the Fe2NiSe4/FeNi3 electrocatalyst can facilitate highly efficient electrochemical splitting of H2S, yielding valuable chemical sulfur and simultaneous production of energy-saving H2 production, offering promising features for potential use as an electrode in technological devices.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/pr12102111/s1, Figure S1: Fe2NiSe4/FeNi3 synthesis process diagram; Figure S2: SEM image of Fe2NiSe4/FeNi3; Figure S3: The survey spectrum of X-ray photoelectron spectroscopy for Fe2NiSe4/FeNi3; Figure S4: Image of electrochemical cell; Figure S5: CVs for (a) Fe2NiSe4/FeNi3, (b) NiSe/NF, (c) FeNi3 foam substrate. (d) The capacitive currents at 0.65 V vs. RHE as a function of scan speed for different materials (△j0 = ja-jc); Figure S6: (a) LSV polarization curves of Fe2NiSe4/FeNi3 initial and after 13.3 h reaction. (b) LSV curve with or without stirrer in the reaction cell; Figure S7: SEM images of Fe2NiSe4/FeNi3 after 13.3 h reaction; Figure S8: Contact angle of different electrode materials for liquid sulfur; Table S1: Comparison of SOR performances of FeNiSe/FeNi3 with other anode oxidation reaction.

Author Contributions

Conceptualization, C.T.; methodology, T.D., N.C. and R.F.; formal analysis, C.T., T.D. and Y.D.; investigation, Y.L., Z.L. and H.G.; data curation, T.D. and L.L.; writing—original draft preparation, T.D.; writing—review and editing, C.T., T.D. and Y.D.; supervision, C.T.; project administration, C.T. All authors have read and agreed to the published version of the manuscript.

Funding

We gratefully acknowledge financial support from the Natural Science Foundation of Sichuan (2022NSFSC0023, 2023NSFSC0112).

Data Availability Statement

Data are available upon request to the authors.

Conflicts of Interest

Authors Tong Ding, Nanheng Cen, Rui Fan, Long Li, Yiping Li and Zongshe Liu were employed by the company PetroChina Southwest Oil & Gas Field Company. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest. The PetroChina Southwest Oil & Gas Field Company had role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Zou, C.; Yang, Z.; He, D.; Wei, Y.; Li, J.; Jia, A.; Chen, J.; Zhao, Q.; Li, Y.; Li, J. Theory, Technology and Prospects of Conventional and Unconventional Natural Gas. Pet. Explor. Dev. 2018, 45, 604–618. [Google Scholar] [CrossRef]
  2. Dan, M.; Yu, S.; Li, Y.; Wei, S.; Xiang, J.; Zhou, Y. Hydrogen Sulfide Conversion: How to Capture Hydrogen and Sulfur by Photocatalysis. J. Photochem. Photobiol. C Photochem. Rev. 2020, 42, 100339. [Google Scholar] [CrossRef]
  3. Sun, M.; Wang, X.; Zhao, Z.; Qiu, J. Review of H2S Selective Oxidation over Carbon-Based Materials at Low Temperature: From Pollutant to Energy Storage Materials. New Carbon. Mater. 2022, 37, 675–694. [Google Scholar] [CrossRef]
  4. Liu, S.; Gu, X.; Yang, S.; Qian, Y. Can Green Hydrogen and Waste Heat Utilization Improve Energy Conservation and Emission Reduction of Coal-Based Cogeneration Processes? J. Clean. Prod. 2023, 389, 136045. [Google Scholar] [CrossRef]
  5. Segawa, Y.; Endo, N.; Shimoda, E.; Yamane, T.; Maeda, T. Pilot-Scale Hydrogen Energy Utilization System Demonstration: A Case Study of a Commercial Building with Supply and Utilization of off-Site Green Hydrogen. Int. J. Hydrogen Energy 2023, 50, S0360319923032731. [Google Scholar] [CrossRef]
  6. Hwang, J.; Maharjan, K.; Cho, H. A Review of Hydrogen Utilization in Power Generation and Transportation Sectors: Achievements and Future Challenges. Int. J. Hydrogen Energy 2023, 48, 28629–28648. [Google Scholar] [CrossRef]
  7. Huang, Y.; Zhu, L.; He, Y.; Wang, Y.; Hao, Q.; Zhu, Y. Carbon Dioxide Utilization Based on Exergoenvironmental Sustainability Assessment: A Case Study of CO2 Hydrogenation to Methanol. Energy 2023, 273, 127219. [Google Scholar] [CrossRef]
  8. Tang, J.; Xu, X.; Tang, T.; Zhong, Y.; Shao, Z. Perovskite-Based Electrocatalysts for Cost-Effective Ultrahigh-Current-Density Water Splitting in Anion Exchange Membrane Electrolyzer Cell. Small Methods 2022, 6, 2201099. [Google Scholar] [CrossRef] [PubMed]
  9. Abdelghafar, F.; Xu, X.; Jiang, S.; Shao, Z. Designing Single-Atom Catalysts toward Improved Alkaline Hydrogen Evolution Reaction. Mater. Rep. Energy 2022, 2, 100144. [Google Scholar] [CrossRef]
  10. Wang, J.; Cui, W.; Liu, Q.; Xing, Z.; Asiri, A.M.; Sun, X. Recent progress in cobalt-based heterogeneous catalysts for electrochemical water splitting. Adv. Mater. 2016, 28, 215–230. [Google Scholar] [CrossRef]
  11. Tang, C.; Zhang, R.; Lu, W.; Wang, Z.; Liu, D.; Hao, S.; Du, G.; Abdullah, M.A.; Sun, X. Energy-saving electrolytic hydrogen generation: Ni2P nanoarray as a high-performance non-noble-metal electrocatalyst. Angew. Chem. Int. Ed. 2017, 56, 842–846. [Google Scholar] [CrossRef] [PubMed]
  12. Kim, K.; Lee, C. Recent Progress in Electrochemical Hydrogen Sulfide Splitting: Strategies for Enabling Sulfur-Tolerant Anodic Reactions. Chem. Eng. J. 2023, 469, 143861. [Google Scholar] [CrossRef]
  13. Kumar, M.; Nagaiah, T.C. Efficient Production of Hydrogen from H 2 S via Electrolysis Using a CoFeS2 Catalyst. J. Mater. Chem. A 2022, 10, 7048–7057. [Google Scholar] [CrossRef]
  14. Kumar, M.; Nagaiah, T.C. Pure Hydrogen and Sulfur Production from H2S by an Electrochemical Approach Using a NiCu–MoS2 Catalyst. J. Mater. Chem. A 2022, 10, 13031–13041. [Google Scholar] [CrossRef]
  15. Xiao, Z.; Lu, C.; Wang, J.; Qian, Y.; Wang, B.; Zhang, Q.; Tang, A.; Yang, H. Bifunctional Co3S4 Nanowires for Robust Sulfion Oxidation and Hydrogen Generation with Low Power Consumption. Adv. Funct. Mater. 2023, 33, 2212183. [Google Scholar] [CrossRef]
  16. Pei, Y.; Cheng, J.; Zhong, H.; Pi, Z.; Zhao, Y.; Jin, F. Sulfide-Oxidation-Assisted Electrochemical Water Splitting for H2 Production on a Bifunctional Cu2S/Nickel Foam Catalyst. Green Chem. 2021, 23, 6975–6983. [Google Scholar] [CrossRef]
  17. Zhang, S.; Zhou, Q.; Shen, Z.; Jin, X.; Zhang, Y.; Shi, M.; Zhou, J.; Liu, J.; Lu, Z.; Zhou, Y. Sulfophobic and Vacancy Design Enables Self-Cleaning Electrodes for Efficient Desulfurization and Concurrent Hydrogen Evolution with Low Energy Consumption. Adv. Funct. Mater. 2021, 31, 2101922. [Google Scholar] [CrossRef]
  18. Li, Y.; Duan, Y.; Zhang, K.; Yu, W. Efficient Anodic Chemical Conversion to Boost Hydrogen Evolution with Low Energy Consumption over Cobalt-Doped Nickel Sulfide Electrocatalyst. Chem. Eng. J. 2022, 433, 134472. [Google Scholar] [CrossRef]
  19. Yu, H.; Wang, W.; Mao, Q.; Deng, K.; Xu, Y.; Wang, Z.; Li, X.; Wang, H.; Wang, L. Electrocatalytic Sulfion Recycling Assisted Energy-Saving Hydrogen Production Using CuCo-Based Nanosheet Arrays. J. Mater. Chem. A 2023, 11, 2218–2224. [Google Scholar] [CrossRef]
  20. Huang, Y.; Lin, L.; Zhang, Y.; Liu, L.; Sa, B.; Lin, J.; Wang, L.; Peng, D.-L.; Xie, Q. Dual-Functional Lithiophilic/Sulfiphilic Binary-Metal Selenide Quantum Dots Toward High-Performance Li–S Full Batteries. Nano-Micro Lett. 2023, 15, 67. [Google Scholar] [CrossRef]
  21. Naaz, F.; Ahmad, T. Ag-Doped WO 3 Nanoplates as Heterogenous Multifunctional Catalyst for Glycerol Acetylation, Electrocatalytic and Enhanced Photocatalytic Hydrogen Production. Langmuir 2023, 39, 9300–9314. [Google Scholar] [CrossRef] [PubMed]
  22. Duan, C.; Tang, C.; Yu, S.; Li, L.; Li, J.; Zhou, Y. Efficient Electrocatalytic Desulfuration and Synchronous Hydrogen Evolution from H2S via Anti-Sulfuretted NiSe Nanowire Array Catalyst. Appl. Catal. B Environ. 2023, 324, 122255. [Google Scholar] [CrossRef]
  23. Yang, J.; Lei, C.; Wang, H.; Yang, B.; Li, Z.; Qiu, M.; Zhuang, X.; Yuan, C.; Lei, L.; Hou, Y. High-Index Faceted Binary-Metal Selenide Nanosheet Arrays as Efficient 3D Electrodes for Alkaline Hydrogen Evolution. Nanoscale 2019, 11, 17571–17578. [Google Scholar] [CrossRef] [PubMed]
  24. Xu, X.; Wang, W.; Zhou, W.; Shao, Z. Recent Advances in Novel Nanostructuring Methods of Perovskite Electrocatalysts for Energy-Related Applications. Small Methods 2018, 2, 1800071. [Google Scholar] [CrossRef]
  25. Yu, T.; Wang, Z.; Tan, K.; He, H.; Yin, S. Novel Hierarchical FeWO4–FeNi3 Heterostructure Nanosheets for Highly Active Water Electrolysis at Large Current Density. Mater. Today Phys. 2023, 35, 101138. [Google Scholar] [CrossRef]
  26. Lambert-Andron, B.; Berodias, G.; Babot, D. Structures magnetiques des composes MFe2Se4 (M= Ti, V, Cr, Fe, Co, Ni). J. Phys. Chem. Solids 1972, 33, 87–94. [Google Scholar] [CrossRef]
  27. Xiao, K.; Zhou, L.; Shao, M.; Wei, M. Fabrication of (Ni,Co)0.85 Se Nanosheet Arrays Derived from Layered Double Hydroxides toward Largely Enhanced Overall Water Splitting. J. Mater. Chem. A 2018, 6, 7585–7591. [Google Scholar] [CrossRef]
  28. Liang, J.; Li, S.; Li, F.; Zhang, L.; Jiang, Y.; Ma, H.; Cheng, K.; Qing, L. Defect Engineering Induces Mo-Regulated Co9Se8/FeNiSe Heterostructures with Selenium Vacancy for Enhanced Electrocatalytic Overall Water Splitting in Alkaline. J. Colloid. Interface Sci. 2024, 655, 296–306. [Google Scholar] [CrossRef]
  29. Xu, R.; Wu, R.; Shi, Y.; Zhang, J.; Zhang, B. Ni3Se2 Nanoforest/Ni Foam as a Hydrophilic, Metallic, and Self-Supported Bifunctional Electrocatalyst for Both H2 and O2 Generations. Nano Energy 2016, 24, 103–110. [Google Scholar] [CrossRef]
Figure 1. (a) XRD image of Fe2NiSe4/FeNi3; (b,c) SEM images of Fe2NiSe4/FeNi3; (d) TEM image of one Fe2NiSe4 nanowire; (e) HRTEM image; and (f) EDS images of Fe2NiSe4/FeNi3.
Figure 1. (a) XRD image of Fe2NiSe4/FeNi3; (b,c) SEM images of Fe2NiSe4/FeNi3; (d) TEM image of one Fe2NiSe4 nanowire; (e) HRTEM image; and (f) EDS images of Fe2NiSe4/FeNi3.
Processes 12 02111 g001
Figure 2. XPS spectra of high-resolution spectrum (a) Ni 2p; (b) Fe 2p; (c) Se 3d for Fe2NiSe4/FeNi3.
Figure 2. XPS spectra of high-resolution spectrum (a) Ni 2p; (b) Fe 2p; (c) Se 3d for Fe2NiSe4/FeNi3.
Processes 12 02111 g002
Figure 3. (a) LSV polarization curves of Fe2NiSe4/FeNi3, NiSe/NF, and bare FeNi3 foam with a scan rate of 5 mv s−1 for sulfide oxidation reaction. (b) The comparison of current densities obtained on various catalysts at different potentials. (c) LSV polarization curves of Fe2NiSe4/FeNi3, NiSe/NF, bare FeNi3 foam, and Pt plate with a scan rate of 5 mv s−1 for hydrogen evolution reaction and (d) LSV polarization curves of SOR and OER of Fe2NiSe4/FeNi3 electrode. (e) Electrochemical impedance spectra of Fe2NiSe4/FeNi3, NiSe/NF, and bare FeNi3 foam, measurement in the electrolyte of 1.0 M NaOH with 1.0 M Na2S. (f) Chronoamperometry curve of Fe2NiSe4/FeNi3 at 0.4 V vs. RHE for 800 min.
Figure 3. (a) LSV polarization curves of Fe2NiSe4/FeNi3, NiSe/NF, and bare FeNi3 foam with a scan rate of 5 mv s−1 for sulfide oxidation reaction. (b) The comparison of current densities obtained on various catalysts at different potentials. (c) LSV polarization curves of Fe2NiSe4/FeNi3, NiSe/NF, bare FeNi3 foam, and Pt plate with a scan rate of 5 mv s−1 for hydrogen evolution reaction and (d) LSV polarization curves of SOR and OER of Fe2NiSe4/FeNi3 electrode. (e) Electrochemical impedance spectra of Fe2NiSe4/FeNi3, NiSe/NF, and bare FeNi3 foam, measurement in the electrolyte of 1.0 M NaOH with 1.0 M Na2S. (f) Chronoamperometry curve of Fe2NiSe4/FeNi3 at 0.4 V vs. RHE for 800 min.
Processes 12 02111 g003
Figure 4. (a) The production rates of H2 in a galvanostatic test at 100 mA cm−2; (b) the faradaic eff-ciencies of H2; (c) the UV-Vis pattern of diluting the electrolyte 20 times with different reaction times; (d) the XRD pattern of the sample obtained (inset: picture of the collected outcome).
Figure 4. (a) The production rates of H2 in a galvanostatic test at 100 mA cm−2; (b) the faradaic eff-ciencies of H2; (c) the UV-Vis pattern of diluting the electrolyte 20 times with different reaction times; (d) the XRD pattern of the sample obtained (inset: picture of the collected outcome).
Processes 12 02111 g004
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

Ding, T.; Cen, N.; Fan, R.; Li, L.; Du, Y.; Tang, C.; Guo, H.; Li, Y.; Liu, Z. Fe2NiSe4 Nanowires Array for Highly Efficient Electrochemical H2S Splitting and Simultaneous Energy-Saving H2 Production. Processes 2024, 12, 2111. https://doi.org/10.3390/pr12102111

AMA Style

Ding T, Cen N, Fan R, Li L, Du Y, Tang C, Guo H, Li Y, Liu Z. Fe2NiSe4 Nanowires Array for Highly Efficient Electrochemical H2S Splitting and Simultaneous Energy-Saving H2 Production. Processes. 2024; 12(10):2111. https://doi.org/10.3390/pr12102111

Chicago/Turabian Style

Ding, Tong, Nanheng Cen, Rui Fan, Long Li, Yonghong Du, Chun Tang, Heng Guo, Yiping Li, and Zongshe Liu. 2024. "Fe2NiSe4 Nanowires Array for Highly Efficient Electrochemical H2S Splitting and Simultaneous Energy-Saving H2 Production" Processes 12, no. 10: 2111. https://doi.org/10.3390/pr12102111

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