Next Article in Journal
Facile Preparation of High-Performance Non-Enzymatic Glucose Sensors Based on Au/CuO Nanocomposites
Previous Article in Journal
Post-Synthesis Ion Beam Sputtering of Pt/CeO2–ZrO2 Catalysts: Correlating Surface Modifications with Light-Off Performance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advances in Electrocatalytic Hydrogen Sulfide Splitting for Sulfur Recovery: From Reaction Mechanisms to Application

1
School of Chemistry and Chemical Engineering, Sichuan University of Arts and Science, Dazhou 635000, China
2
Basalt Fiber and Composite Key Laboratory of Sichuan Province, Sichuan University of Arts and Science, Dazhou 635000, China
3
Dazhou Key Laboratory of Advanced Technology for Fiber Materials, Sichuan University of Arts and Science, Dazhou 635000, China
4
Dazhou Ecological and Environmental Science Research Institute, Dazhou 635000, China
*
Author to whom correspondence should be addressed.
Catalysts 2025, 15(11), 1019; https://doi.org/10.3390/catal15111019
Submission received: 18 September 2025 / Revised: 19 October 2025 / Accepted: 28 October 2025 / Published: 30 October 2025

Abstract

Hydrogen sulfide (H2S), a highly toxic gas, is mainly sourced from petroleum refining, natural gas purification, and coal chemical processes. It poses significant risks to human health, causes environmental pollution, and accelerates equipment corrosion. Recent studies have demonstrated that electrochemical coupling systems offer an efficient, sustainable, and cost-effective strategy for removing sulfur-containing gaseous pollutants. These systems enable the conversion of H2S into recoverable sulfur under mild conditions, while simultaneously harnessing the chemical energy of H2S to drive the production of higher-value products (H2, HCOOH, CH4, CO, H2O2, etc.). Therefore, electrochemical systems for sulfur recovery have received increasing attention. This review highlights the significance of electrochemical recovery of sulfur from H2S. It summarizes the reaction pathways and mechanisms involved in anodic sulfur oxidation, critically analyzes and discusses methods for detecting sulfur oxidation products, and summarizes the latest advances in sulfur oxidation reaction (SOR) anode materials and various electrochemical coupling systems. The aim is to enhance the fundamental understanding of electrochemical sulfur recovery and to provide insights for the design of novel SOR electrodes and integrated electrochemical coupling systems.

1. Introduction

With the sustained global growth of global energy demand, the extraction and processing of fossil fuels (oil, natural gas, and coal) and the development of biomass energy (biogas) have expanded rapidly, leading to massive emissions of hydrogen sulfide (H2S) [1,2,3]. Statistics show that the global annual H2S emissions from industries such as petroleum refining, natural gas purification, and coal chemical engineering exceed 30 million tons [3,4], presenting a severe environmental burden. Notably, in the purification process of certain high-sulfur natural gas reserves, the volume fraction of H2S can even surpass 40% [5]. H2S poses multifaceted and severe threats to human health, industrial safety, and ecological stability: inhalation of air with H2S concentrations exceeding 300 ppb can cause immediate respiratory arrest [6]; its inherent corrosivity degrades pipelines and equipment, substantially increasing maintenance costs for sulfur-containing gas transportation infrastructure [7,8,9]; furthermore, its release into the atmosphere participates in photochemical reactions, exacerbating acid rain and smog events, which in turn disrupt terrestrial and aquatic ecosystems [10].
As a globally strategic resource, sulfur is indispensable to the chemical, agriculture, and pharmaceutical industries, with a global annual demand of over 70 million tons [4,11]. Traditional H2S treatment technologies, typified by the Claus process, achieve sulfur recovery but suffer from inherent limitations: severe harsh operating conditions (800–1400 °C), inevitable SO2 by-products, and low efficiency for low-concentration H2S (<5 vol%) [12]. In recent years, electrocatalytic oxidation technology has emerged as a promising alternative for H2S decomposition and sulfur recovery, featuring mild conditions (ambient temperature and pressure), high sulfur selectivity, and reduced energy consumption [2,3]. Under an external electric field, H2S is oxidized to sulfur at the anode, while value-added reduction reactions (hydrogen evolution, CO2-to-CO conversion) occur at the cathode, realizing the integration of “waste detoxification-sulfur recovery-energy/chemical production” [13,14].
Owing to the unique advantages of electrocatalytic sulfur recovery technology in H2S treatment—including mild reaction conditions, high sulfur selectivity, and the capability to couple with the production of high-value products (e.g., H2, CO)-research in this field has grown increasingly abundant, now covering catalyst design, reaction mechanism analysis, and practical wastewater treatment applications [3,11]. Conducting a thematic review here holds great academic and practical significance to systematically organize the research context, clarify core breakthroughs and unresolved issues, and provide clear exploration guidelines for subsequent researchers. However, relevant reviews remain relatively scarce, and existing research summaries generally lack a systematic elaboration of the “reaction mechanism-catalytic materials-oxidation products-coupling systems” framework: for instance, Yu et al. conducted in-depth analyses of the core mechanism of sulfide oxidation reaction (SOR, e.g., the stepwise oxidation pathway of S2−→Sn2−→S8) and electrode material design, and clarified the effect of d-band center modulation on sulfur adsorption/desorption equilibrium via density functional theory (DFT) calculations, but failed to systematically elaborate on the formation principles and detection methods of sulfur oxidation products and completely omitted research on emerging coupling systems such as SOR-CO2RR [3]; additionally, Shao et al. carried out extensive research on coupling systems for electrochemical sulfur oxidation, optimized operating parameters for practical wastewater treatment, and verified the technology’s applicability, yet they lacked systematic analysis of the catalytic process on the electrode surface and failed to establish a direct correlation between catalytic material structures (e.g., defects, crystal facets) and catalytic performance (e.g., overpotential, selectivity), making it difficult to support the targeted design of subsequent catalysts [3,11,12].
This review summarizes progress in electrocatalytic H2S splitting for sulfur recovery, structured by the logic chain “reaction mechanism→key components (electrode materials, reactor systems)→product analysis→technical significance”—a framework addressing gaps in existing literature. Unlike prior fragmented reviews, it broadens scope by integrating emerging electrochemical coupling systems and tackling mass transfer limitations in low-concentration H2S, a key industrial issue long neglected; deepens mechanistic analysis by linking Sn2− polymerization kinetics and S8 desorption energy barriers to solutions for electrode passivation, beyond superficial pathway descriptions; and sharpens focus on catalysts/processes by classifying electrodes by function (e.g., high-efficiency S8 production, controllable Sn2− generation) and clarifying structure-performance relationships, moving past unstructured material lists. It further identifies core challenges (sulfur-induced passivation, dilute H2S mass transfer) and proposes optimization directions, aiming to bridge fundamental research and industrial application.

2. The Significance of Electrochemical Recovery of Sulfur

2.1. Environmental Significance

H2S is widely recognized as an atmospheric and water pollutant, and its harm to the environment is characterized by multi-dimensionality and long-term effects [15,16]. Given its strong toxicity and widespread emissions, H2S poses severe threats to ecological stability: In the atmosphere, H2S readily reacts with oxygen, ozone, and other substances to form SO2, sulfate, and other compounds [17]. Among these products, SO2 is one of the primary precursors of acid rain, contributing to soil acidification, forest degradation, and eutrophication of water bodies—ultimately disrupting the ecological balance [18]. Sulfate aerosols, on the other hand, reduce atmospheric visibility and exacerbate smog pollution [19,20]. In aquatic environments, H2S dissolves to form hydrosulfuric acid, which lowers the pH of water bodies, inhibits the respiratory function of aquatic organisms, causes mass mortality of fish, algae, and other biota, and thus disrupts aquatic ecosystems [21]. Additionally, H2S can infiltrate and contaminate soil, impairing crop growth and reducing the quality of agricultural products [22,23].
Electrochemical sulfur recovery technology, relying on electrocatalytic oxidation, converts sulfides in H2S directly into sulfur while avoiding the generation of secondary pollutants such as SO2 that are inherent in traditional treatment methods (e.g., the Claus process) [1,24]. Moreover, this technology operates under mild ambient temperature and pressure conditions, eliminating the requirement for high temperatures or pressures. This not only significantly reduces process energy consumption and carbon emissions but also aligns with the “carbon peak and carbon neutrality” goals. For industrial wastewater containing H2S, this electrochemical sulfur recovery technology can further enable in situ treatment, avoiding the risk of leakage during wastewater transportation and thereby reducing pollution to aquatic environments. Therefore, electrochemical sulfur recovery technology holds great practical significance in controlling H2S pollution and protecting the ecological environment.

2.2. Economic Benefits

From the perspective of resource utilization, the sulfur element in H2S is a valuable chemical raw material, and the recovery of sulfur can bring significant economic benefits [1,25,26]. Sulfur, as a crucial basic chemical product, is widely applied in the production of sulfuric acid, fertilizers (such as superphosphate and ammonium sulfate), rubber vulcanizing agents, and pharmaceutical intermediates [27]. The global demand for sulfur is stable, and the global sulfur price remained at USD 150–200 per ton in 2024, as reported by IndexBox. Moreover, with the rapid development of the energy industry, the demand for sulfuric acid in the production of lithium battery materials is increasing, further boosting the market value of sulfur [28].
Traditional H2S treatment technologies, typified by the Claus process, can recover sulfur but suffer from high equipment investment, large land occupation, elevated operating costs, and poor economic efficiency when treating low-concentration H2S (volume fraction < 5 vol%). Consequently, substantial amounts of low-concentration H2S are directly discharged or incinerated without sulfur recovery, resulting in the waste of sulfur resources. In contrast, electrochemical sulfur recovery technology features compact equipment, simple operation, and strong adaptability to H2S concentrations [26]. It can not only handle high-concentration H2S but also efficiently recover sulfur from low-concentration H2S, greatly improving the utilization rate of sulfur resources. Additionally, the sulfide oxidation reaction (SOR) as the core anodic reaction at the anode can be coupled with specific reduction reactions at the cathode of the electrochemical system, namely the hydrogen evolution reaction (HER) and carbon dioxide reduction reaction (CO2RR). This coupling achieves the synergistic benefits of “sulfur recovery + energy/chemical production”, thereby further enhancing the economic value of the technology [11,29].

3. SOR Mechanism

In the SOR, the oxidation products of S2− include various short-chain polysulfides (Sn2−, 2 ≤ n ≤ 6), S2O32−, SO32−, SO42−, etc.) [3,13,30]. Depending on the subsequent reaction pathways of the SOR intermediate products and the route of sulfur recovery, the SOR pathways can be classified into two types (Figure 1): (1) direct oxidation for sulfur recovery; (2) indirect oxidation for sulfur recovery [31,32]. SOR for sulfur recovery is achieved by successive oxidation reactions of S2− [31]. S2− is first gradually oxidized to form Sn2−, and then Sn2− is further oxidized to eventually generate the high-value target product S8 [29,32]. The entire reaction does not require the introduction of additional media, and the recovery of S8 can be accomplished solely by successive oxidation steps [29]. Indirect oxidation for sulfur recovery differs from the direct oxidation pathway. This process requires the use of external oxidants (such as Fe3+, I3, EDTA-Fe3+, etc.) or reductants (such as S2−, S2O32−, etc.) as reaction media [29,33]. Through the oxidation/reduction reactions between the media and S2− or SOR intermediate products, the intermediate products are converted into sulfur. The core of this process is to rely on exogenous active substances to drive the conversion of S2− or SOR intermediate products into S8.

3.1. Direct Oxidation Pathway

In the direct oxidation pathway, sulfide directly adsorbs onto the active sites of the anode surface and transfers electrons directly to the electrode to undergo oxidation reactions and form sulfur [7,10,34]. The direct oxidation pathway follows a stepwise oxidation logic of “S2−→Sn2−→sulfur” (* + S2−→*S2−→*S22−→*S32−→*S42−→*S82−→* + S8, * denotes the active sites on the catalyst surface), with Sn2− as the main intermediate product, and the generation and transformation of each intermediate have clear potential dependencies [35]. The process can include three key reactions. (1) S2− is gradually oxidized to form Sn2−. Under alkaline conditions, HS/S2− first adsorbs onto the active sites of the anode and gradually polymerizes to Sn2− through single-electron or multi-electron transfer. For example, Xiao et al.’s research shows that on the surface of Co3S4 nanowire catalysts, S2− is first oxidized to S22− (adsorption energy −0.8 to −1.2 eV) and then further converted to intermediate products such as S32− and S42−, demonstrating the dynamic generation process of Sn2− [2]. (2) Sn2− is oxidized to form sulfur. Sn2− is the key intermediate in the direct oxidation pathway, and its further oxidation is the rate-controlling step of the reaction. Density functional theory (DFT) calculations show that on the Co3S4 (100) crystal plane, the energy barrier for the oxidation of S32− to S42− is only 1.40 eV, much lower than 1.65 eV on the (110) crystal plane and 1.85 eV on the (311) crystal plane. This low energy barrier ensures the efficient conversion of Sn2−; when the potential rises from 0.4 to 0.6 V (vs RHE), S42− and other intermediates further polymerize through S-S bonds and eventually form S8 [2]. (3) In addition, the desorption ability of sulfur is the core bottleneck of the direct oxidation pathway. High desorption ability of sulfur can lead to its further oxidation to by-products such as S2O32−, SO32−, and SO42−, resulting in excessive oxidation of sulfur, or sulfur adsorbing on the electrode surface and causing electrode passivation [3,30]. The adsorption energy of S8 on the Co3S4 (311) crystal plane is only 2.52 eV [2], much lower than that of traditional catalysts such as Ni (4.27 eV) and Pt (4.39 eV) [35], effectively avoiding the passivation caused by the deposition of S8 on the electrode surface. NiS2 further accelerates the desorption of S8 through a “sulfur-repellent” surface design (contact angle > 90°), ensuring the continuous progress of the reaction [35].
To address the electrode passivation issue caused by sulfur, a novel SOR pathway has been proposed [29]. The core mechanism of this pathway is as follows: Firstly, HS/S2 are oxidized to Sn2− on the electrode surface (Equation (1)); subsequently, Sn2− is treated with reagents such as CO2, HCl, or H2SO4 (Equation (2)), ultimately achieving efficient recovery of sulfur. This strategy successfully avoided the key problem of electrode passivation in the direct oxidation recovery process, providing a new direction for the resourceful treatment of sulfur-containing wastewater [29]. In the related research field, Huang et al. were the first to adopt a similar approach [36]. They prepared a ZIS/NiFeS composite through electrochemical deposition, and this electrode could selectively oxidize S2− to Sn2−, thereby enabling the collection of S8. These research results have verified the effectiveness and application potential of this strategy in the treatment of sulfur-containing wastewater from different experimental systems, providing crucial support for its further promotion [3,29]. Additionally, Chen et al.’s research has deeply revealed the multi-path characteristics of the oxidation of HS/S2− to Sn2−, and clarified the regulatory effect of HS/S2−concentration on the oxidation products [30]. When the S2− concentration is low (100 mg/L), it is more likely to be deeply oxidized to SO42−, resulting in a significant decrease in the recovery rate of elemental sulfur. When the S2− concentration is increased to 500 mg/L, the oxidation products are mainly Sn2− (such as S22− and S32−), which are more conducive to the subsequent recovery of sulfur. Detailly, under low concentration conditions, HS/S2− is first oxidized to S2O32−, and S2O32− is further oxidized to SO42−; under high concentration conditions, HS/S2 also initially generates S2O32−, but subsequently S2O32− reacts with the excess S2− in the system to form sulfur (S) (Equation (3)), and the generated S then complexes with S2− to form Sn2− (Equation (4)), ultimately creating favorable conditions for the recovery of sulfur.
n S 2 + 2 n 2 h + S n 2 2 n 4
n S 2 + 2 n 2 h + S n 2 2 n 4
S 2 O 3 2 + 2 S 2 + 3 H 2 O 4 S + 6 O H
S + 2 n H S + O H S n 2 + H 2 O
The reaction efficiency of the direct oxidation pathway mainly depends on the number of active sites on the electrode surface, the adsorption capacity of HS/S2−, and the electron transfer efficiency. Generally, electrode materials with high specific surface area and rich defect structures can provide more active sites and promote the adsorption of reactants and electron transfer, thereby enhancing the efficiency of the direct oxidation reaction. In addition, the pH value of the electrolyte significantly affects the direct oxidation pathway [30]. Under alkaline conditions, the dissociation degree of H2S is high, and the concentration of HS/S2− is large, which is more conducive to the direct oxidation reaction; while under acidic conditions, H2S mainly exists in molecular form, with weak adsorption capacity, and the direct oxidation reaction rate is lower [37].

3.2. Indirect Oxidation Pathway

The indirect oxidation pathway of SOR, dominated by redox mediators, refers to the oxidation of HS/S2− being accomplished through the oxidation state intermediates generated on the electrode surface as electron transfer mediators [1,38,39]. That is, the electrode first oxidizes the mediator to form an oxidizing intermediate, which then chemically reacts with HS/S2− to oxidize it to elemental sulfur, while the intermediate itself is reduced, forming a cycle of “electrode oxidation of mediator-intermediate oxidation of HS/S2−-intermediate regeneration” (Equations (5)–(10)) [3,40,41]. Common indirect oxidation intermediates include high-valent ions such as Fe3+, Co3+, Ni3+, I3 [29]. Take the indirect oxidation process with Fe3+ as the intermediate as an example. Intermediate generation: At the anode, Fe2+ is oxidized to Fe3+ (Equation (7)). Intermediate oxidation of H2S: The generated Fe3+ enters the electrolyte and undergoes a chemical reaction with H2S/HS/S2−, oxidizing them to sulfur, while Fe3+ is reduced to Fe2+ (Equation (8)). Intermediate regeneration: The reduced Fe2+ re-diffuses to the anode surface and is oxidized back to Fe3+, completing the recycling of the intermediate. The overall reaction equation is the same as that of the direct oxidation pathway (Equation (11)).
3 I + 2 h + I 3
I 3 + S 2 3 I + S
F e 2 + + h + F e 3 +
2 F e 3 + + S 2 2 F e 2 + + S
E D T A F e 2 + + h + E D T A F e 3 +
2 E D T A F e 3 + + S 2 2 E D T A F e 2 + + S
8 S 2 8 e S 8
The advantage of the indirect oxidation pathway lies in the existence of intermediates can lower the activation energy of the SOR and increase the reaction rate [29]. Especially for electrode materials with weak electron transfer ability, the indirect oxidation pathway can significantly improve the reaction performance. In addition, the concentration, redox potential, and stability of the intermediates have a significant impact on the efficiency of indirect oxidation [29]. For example, the redox potential of Fe3+ (0.77 V vs. SHE) is moderate, which can effectively oxidize sulfide without easily causing over-oxidation to form SO42−; while the redox potential of Co3+ is relatively high (1.81 V vs. SHE), although it has strong oxidation ability, it is prone to over-oxidation of sulfide, reducing the selectivity of sulfur [42].

4. Analysis of Sulfur Oxidation Products

The products of SOR are complex and diverse, encompassing not only the target product S8 but also a range of intermediates and by-products such as Sn2−, S2O32−, SO32−, and SO42−. Accurate analysis and identification of the types, contents, and existing forms of these sulfur oxidation products hold crucial significance for optimizing reaction conditions, improving SOR selectivity, and elucidating the underlying reaction mechanism. At present, the main analytical methods for SOR products are XRD, Raman, UV-vis spectrometer (UV-vis), and gas chromatography-mass spectrometry (Figure 2) [3,43].
The SOR value-added product S8 is usually confirmed by XRD and Raman analysis. For example, Xiao et al. analyzed the SOR product by XRD and found that the collected product powder corresponded to sulfur (S8, PDF #78-1888) [2]. Chen et al. analyzed the light yellow powder collected from SOR by XRD and found that the product powder corresponded to elemental sulfur (S8, PDF #74-1465) [30]. In addition, Raman spectroscopy is also an effective method for analyzing sulfur. Xiao et al. found that when NiMoN@C/NF was used as the SOR anode, a new characteristic peak of S8 appeared at 461 cm−1 when the SOR voltage rose to 0.4 V (vs. RHE), and three characteristic peaks of S8 at 146, 216, and 461 cm−1 were observed when the applied voltage was greater than 0.45 V (vs. RHE) [2]. Similarly, He et al. constructed an in situ electrochemical Raman system using a self-made in situ Raman device [44]. During the potentiation test at a constant current of 50 mA/cm2, a laser wavelength of 532 nm was used for testing every minute, and the absorption peak of S8 was detected at a Raman wavelength of 220 cm−1.
To gain a deeper understanding of the formation pathway of S8, it is necessary to conduct an in-depth analysis of the intermediate products of SOR. In the direct oxidation pathway (* + S2−→*S2−→*S22−→*S32−→*S42−→*S82−→* + S8), the main intermediate products are Sn2−. The adsorbed intermediate products * Sn2− are mainly detected by in situ Raman spectroscopy. He et al. used in situ Raman spectroscopy with an excitation wavelength of 532 nm and detected Sn2− at 450~550 nm, directly proving the pathway of SOR to produce S8 [44]. However, for free Sn2−, UV-vis spectrophotometry is currently used for detection. Chen et al. used a UV-vis (UV-2100, Shanghai Mapada Instruments Co., Ltd., Shanghai, China) and detected Sn2− (2≤ n ≤4) at absorption wavelengths of 300 and 370 nm [2]. However, UV-vis cannot determine the specific value of n in Sn2−. Zheng et al.’s research shows that GC-MS can be used to accurately analyze Sn2−. This study used trifluoromethanesulfonic acid methyl ester (CF3SO3CH3) as a methylation reagent to react with Sn2− (Equation (12)) [45]. CF3SO3CH3 can convert Sn2− into (CH3)2Sn. The reaction principle is that the methyl group (-CH3) in CF3SO3CH3 undergoes methylation with Sn2−, and then the generated (CH3)2Sn is determined by GC-MS [43]. For example, Chen et al. used this method to precisely distinguish between two S2− oxidation products, namely (CH3)2S2 (m/z = 93.9910) and (CH3)2S3 (m/z = 125.9631) [30]. In addition to the above core species, the analysis methods for other sulfur species in the SOR system have been well established: S2− was tested by the spectrophotometric method using p-amino dimethylaniline. S2O32− is commonly determined by titration techniques such as iodometry, while the concentrations of SO32− and SO42− in solution were analyzed using ion chromatography [30,46].
S n 2 + 2 C F 3 S O 3 C H 3 C H 3 S n + 2 C F 3 S O 3

5. Electrodes for Recovering Sulfur

In the SOR process, the electrodes serve as the core carriers for electron transfer. Their material composition, microstructure, and surface properties directly determine the oxidation pathway of S2−, reaction kinetics, and product selectivity [3]. Based on differences in the target products of SOR and reaction mechanisms, electrodes for sulfur recovery can be categorized into three main types (Table 1): (1) high-efficiency sulfur-producing electrodes, (2) controllable S2−-producing electrodes, and (3) medium circulation-driven indirect oxidation electrodes.

5.1. Direct Oxidation Electrodes

Direct oxidation-type electrodes can efficiently catalyze S2− to directly generate S8 through one or multiple electron transfer steps [3,29]. Due to the advantages of short reaction paths, low energy loss, and high product selectivity, they have become the primary focus of current research on electrodes for elemental sulfur recovery [33,59]. The key to their design lies in constructing electrode surfaces with high active site density, strong S2− adsorption capacity, and excellent electronic conductivity, while suppressing the formation of by-products such as S2O32− and SO42− [3,60,61,62]. Currently, the more mature types mainly include transition metal sulfide (TMS) electrodes, alloy and high-entropy alloy (HEA) electrodes, metal-carbon composite electrodes, and noble metal sulfide electrodes [3,33]. Among them, TMS electrodes follow the “adsorption-electron transfer-polymerization-desorption” path (S2−→*S→*S2→…→*S8→S8), and by regulating the strength of the metal-sulfur (M-S) bond to inhibit passivation [63,64]. For example, NiS2 has a “sulfur-repellent” surface that can repel S8 deposition, has an overpotential of only 0.41 V at 10 mA/cm2 in 0.05 mol/L Na2S electrolyte, a sulfur recovery rate of 92%, and no obvious passivation after 100 h of operation [35]. Ni3S2, after Co doping to regulate the d-band center, can shorten the oxidation path, and the cell voltage drops to 0.80 V at 50 mA/cm2 [65]. Alloy and high-entropy alloy electrodes regulate the d-band center through alloying to balance S2− adsorption and S8 desorption [66]. High-entropy alloys can also enhance structural stability by high configurational entropy [67,68,69,70,71]. For example, Pd8Co2/C has an overpotential of −0.35 V (vs. Ag/AgCl) in 1 mol/L Na2S + 3 mol/L KOH at 100 mA/cm2, enabling the maximum power density of direct alkaline sulfur fuel cells to reach 46.82 mW/cm2 (26% higher than Pd/C) [49], and Pd9Ir1/C has a power density of 33.98 mW/cm2 at 70 °C [72]. The FeCoNiCuMn HEA and Mo2C composite electrode, due to the synergy of multiple elements and the low SOR energy barrier of the Mo2C (102) crystal plane, has an overpotential of 0.382 V at 100 mA/cm2 and is suitable for both HER and SOR [55]. Metal-carbon composite electrodes rely on carbon substrates to promote electron conduction and mass diffusion and utilize the charge transfer effect at the metal-carbon interface to optimize the S2− adsorption energy barrier [3,73,74]. For example, CoNi@NGs has an initial SOR potential of 0.25 V and a current density twice that of Pt/C in 1 mol/L Na2S + 1 mol/L KOH [47].

5.2. Controllable Sn2− Producing Electrodes

Controllable Sn2−-producing electrodes aim to selectively catalyze the oxidation of S2−/HS to specific Sn2− as the core objective [3,30]. By precisely regulating the reaction pathway to avoid over-oxidation to by-products such as SO32− and SO42−, they provide controllable intermediates for subsequent stepwise conversion to elemental sulfur [3,35,36,44]. The key to their design lies in matching the energy barrier requirements for the stepwise oxidation of S2− and enhancing product selectivity [3]. Typical electrode materials include transition metal sulfides, doped metal oxides, and high-entropy sulfides [3]. For example, Cu2S/NF electrodes utilize the Cu+/Cu2+ redox couple to modulate electron transfer, limiting S2− oxidation to the Sn2− stage by controlling the electron transfer rate [52]; Sn4+-doped α-Fe2O3 introduces defect sites via doping while leveraging the Fe3+/Fe2+ redox activity, enhancing S2− adsorption and inhibiting deep oxidation [53]; high-entropy sulfide CuCoNiMnCrSx/NF achieves precise regulation of Sn2− chain length (mainly n = 2~4) through the synergistic effect of multiple elements, which optimizes the d-band center to balance S2− adsorption and Sn2− desorption [33]; B2S3 electrodes rely on the valence change of Bi3+/Bi2+ to adjust catalytic activity, and preferentially generate short-chain Sn2− under mild potentials (0.2~0.4 V vs. RHE) [54]. Additionally, electrodes like NF/Mo-Co(OH)2 use Co(OH)2’s surface hydroxyl groups to selectively adsorb S2−, and the intercalated structure restricts the growth of polysulfide chains, further ensuring the selectivity of Sn2− [56]. Such electrodes are particularly suitable for sulfur-containing wastewater treatment systems, as the generated Sn2− can be efficiently converted to S8 by adding acids (e.g., HCl, H2SO4) in subsequent processes, avoiding electrode passivation caused by direct S8 deposition, and their performance can be further optimized by adjusting electrolyte pH (alkaline conditions are more favorable for maintaining Sn2− stability and preventing disproportionation) [30].

5.3. Medium Circulation-Driven Indirect Oxidation Electrodes

Indirect oxidation electrodes do not directly react with S2− but instead achieve the conversion of S2− to S8 through catalyzing the redox cycle of redox mediators [40,41,58]. This system has the advantages of easy product separation and less electrode passivation, making it suitable for treating sulfide wastewater with high concentration and complex composition [3]. According to the type of redox mediators, it can be classified into chemical mediators, indirect oxidation electrodes, and biological mediators indirect oxidation electrodes [29,75,76].
Indirect oxidation systems with chemical redox mediators typically employ transition metal ions (such as Fe3+, EDTA-Fe3+, I3, etc.) or organic compounds (such as quinones) as redox mediators [3]. The role of the electrode is to oxidize the reduced state mediator (such as Fe2+, EDTA-Fe2+, I, etc.) to the oxidized state, and the oxidized state mediator then reacts with S2− in the solution to form S8. For example, in the Fe3+/Fe2+ mediator system, the titanium-based lead dioxide (Ti/PbO2) electrode, due to its high oxygen overpotential, can efficiently catalyze the oxidation of Fe2+ to Fe3+. At a current density of 20 mA/cm2, the oxidation rate of Fe2+ reaches 99%, and the generated Fe3+ reacts with S2− to form S8, with a total sulfur recovery rate exceeding 90% [51,77,78]. In addition, organic quinone mediators (such as anthraquinone-2-sulfonic acid) have the characteristic of adjustable redox potential. After modification of the glassy carbon electrode, it can catalyze the oxidation of quinones, indirectly achieving the mild oxidation of S2−, with a selectivity for S8 reaching 92% [1,76,79].
Biological mediator indirect oxidation electrodes use microorganisms as redox mediators [75,80]. Through the metabolic activities of microorganisms, S2− is oxidized to S8, and the electrode acts as an electron acceptor for the microorganisms, facilitating the electron transfer in the metabolic process. Such electrodes are typically biofilm electrodes, composed of a conductive substrate (such as carbon felt or graphite rods) and a microbial film. For instance, a carbon felt bioelectrode with Thiobacillus thiooxidans attached, under the catalytic action of microorganisms, S2− is first oxidized to Sn2− and then further converted to S8 [75]. At a low voltage (0.2 V vs. RHE), the sulfur recovery rate reaches 88%, and the energy consumption is only 1/3 of that of chemical oxidation methods [75]. In addition, microbial fuel cell-type indirect oxidation electrodes can achieve the synergy of energy production and sulfur recovery [81]. While the electrode catalyzes the microbial metabolism to generate electricity, it indirectly promotes the oxidation of S2− to S8. For instance, a carbon paper electrode with Shewanella oneidensis as the microorganism can achieve an S8 production at a power density of 0.66 mW/cm2 [82]. The key challenge in the bio-mediated system lies in the stability of the microorganisms. By using immobilization techniques (such as sodium alginate entrapment), the adhesion strength of the microorganisms on the electrode surface can be enhanced, thereby extending the service life of the electrode [83].

6. Electrochemical Coupling System for SOR

SOR is not only the core process for recovering elemental sulfur, but also its accompanying electron transfer characteristics and oxidation capacity endow it with great potential for coupling with fields such as energy production and pollution control [29]. By constructing a “sulfur oxidation-targeted function” collaborative system, multiple goals such as “resource recovery-energy output-pollution removal” can be achieved, significantly enhancing the comprehensive value of the sulfur oxidation process [2]. Currently, the mainstream research systems mainly fall into two categories (Table 2): sulfur oxidation coupled with energy substance production and sulfur oxidation coupled with pollutant treatment.

6.1. SOR Coupled with the Production of Energy Substances

Electrocatalytic sulfur oxidation coupled with energy production takes the electrocatalytic system as the core. It releases electrons through the electrocatalytic oxidation of S2− at the anode and combines with the reduction reaction at the cathode (such as oxygen reduction, hydrogen evolution, and CO2 reduction) to achieve the targeted production of clean energy (H2, H2O2, CH4, CO, CHOOH, etc.) [29,30,33]. The key to this system lies in optimizing the activity and selectivity of the anode electrocatalyst, reducing the reaction overpotential, and matching efficient cathode reactions to enhance energy conversion efficiency [29]. Currently, mature coupling modes include electrocatalytic SOR coupled with ORR, electrocatalytic SOR coupled with HER, and electrocatalytic SOR coupled with CO2RR, as shown in Figure 3.

6.1.1. SOR Coupled with ORR

There exists a thermodynamic driving force for the spontaneous reaction between S2− and O2, which provides a thermodynamic foundation for the efficient coupling of the electrocatalytic SOR and ORR [96]. From the perspective of thermodynamic matching, the potential intervals and reaction energy barriers of the two reactions are highly consistent [30]. From a kinetic perspective, by rationally designing the catalyst and reaction system, the electron transfer rate of SOR and the reaction path of ORR can be effectively regulated to achieve a coordinated alignment of the reaction rates at both electrodes, thereby ensuring the stable operation of the entire coupled system [30,38,40]. At the cathode, ORR mainly occurs through two reaction pathways: 4e-ORR and 2e-ORR [30,97]. The standard reduction potential of the 2e-ORR pathway is 0.68 V (vs. RHE), which is much lower than that of the 4e-ORR pathway at 1.23 V (vs. RHE) [98,99,100]. This indicates that the 2e-ORR pathway has a lower reaction energy barrier thermodynamically and is more likely to occur preferentially [30]. Meanwhile, the product of 2e-ORR is H2O2, an environmentally friendly oxidant with strong oxidation capacity and without secondary pollution after reaction [101,102]. It has shown broad application in various fields such as medical disinfection, advanced oxidation treatment of environmental pollutants, energy storage (such as H2O2 fuel cells), and fine chemical synthesis [103,104]. This makes the SOR-2e-ORR coupled system have dual significance in environmental governance and the preparation of high-value chemicals [38]. Typically, Zong et al. constructed a SOR-2e-ORR system (Figure 3a), combining the AQ/H2AQ and I/I3 redox couples, which is based on an unbiased solar photoelectrochemical cell. The cell uses a Nafion membrane to separate the anodic and cathodic compartments, confining the two redox couples, respectively, to avoid backward reactions. This system can utilize solar energy to oxidize toxic H2S to elemental sulfur through the I/I3 redox couple and O2 to valuable H2O2 via the AQ/H2AQ redox couple, with both processes achieving high Faradaic efficiencies [38].

6.1.2. SOR Coupled with HER

The electrocatalytic SOR-assisted H2 production system replaces the traditional oxygen evolution reaction (OER) in water electrolysis with the electrocatalytic oxidation of S2− at the anode, taking advantage of the fact that the oxidation potential of S2− (−0.48 V vs. RHE) is much lower than that of OER (1.23 V vs. RHE), significantly reducing the cell voltage and energy consumption [2,3,44]. The core of this system is to develop anode electrocatalysts with both high activity and stability while matching them with efficient hydrogen evolution cathodes to achieve simultaneous optimization of S2− oxidation and H2 production. In an acidic system, an electrolytic system with defective CoCd(x:y)Sn as the anode and Pt as the cathode has a cell voltage of only 0.25 V, reducing energy consumption, with a high efficiency S2− recovery and a stable H2 production with an achieved 98% H2 Faradaic efficiency with remarkable stability of 120 h [105]. Under alkaline conditions, an electrolytic system with lamellar stacking carbon-confined nickel hybrid nanosheets (TLNi@C HNSs) as bifunctional electrodes (both anode and cathode) has a cell voltage of only 0.91 V at an industrial current density of 1 A cm−2, reducing energy consumption (electricity expense of 2.1 kWh h−1 m−2), with efficient S2− conversion to sulfur (achieving 3.83 kg h−1 m−2 S8 productivity and 69% S8 yield) and stable H2 production, with remarkable stability of over 300 h [106]. Additionally, by constructing a flow-type electrolytic cell system, mass transfer of the electrolyte can be enhanced, reducing the deposition and passivation of S8 on the electrode surface. Zhu et al. constructed a flow-type electrolytic cell system based on anion exchange membranes [106]. By continuously supplying an alkaline electrolyte containing S2− (1.0 M KOH + Na2S) and promptly removing the products on the electrode surface, the mass transfer efficiency of the electrolyte was effectively enhanced, and the deposition and passivation of S8 on the electrode surface were suppressed. This system not only achieved a high current density of 200 mA cm−2 but also reached an industrial-level current density of 1 A cm−2, while significantly reducing the energy consumption per unit hydrogen production.

6.1.3. SOR Coupled with CO2RR

The electrocatalytic SOR-CO2RR coupling system utilizes the electrons generated from the oxidation of S2− at the anode to drive the reduction of CO2 at the cathode to produce carbon-based fuels such as CO, CH4, or CHOOH, achieving the triple goals of “sulfur recovery-carbon emission reduction-energy production” [13,58]. Core to this system is the design of high-performance, low-cost catalysts for both half-reactions: for cathodic CO2RR, non-precious metal catalysts such as graphene-encapsulated ZnO (ZnO@G) and highly dispersed p-Bi nanosheets (p-Bi NSs) have shown excellent selectivity—ZnO@G achieves a Faradaic efficiency for CO of up to 83% and a turnover frequency of 0.024 s−1 at −0.813 V vs. RHE, while p-Bi NSs maintains over 90% FE for formate across a wide potential range (−0.6~1.1 V vs. RHE) [13,58]; for anodic SOR, graphene-modified graphite carbon sheets (G/GCS) and porous Co-S nanosheets (Co-S NSs) exhibit high activity, with G/GCS enabling 100% FE for EDTA-Fe2+ oxidation and Co-S NSs achieving an ultra-low SOR onset potential of 0.2 V vs. RHE [13,58]. The technical challenge of this system lies in matching the electron transfer rates of the anodic S2− oxidation and cathodic CO2 reduction while suppressing side reactions such as HER at the cathode [57,58]. To address this, researchers have optimized reactor configurations and redox mediators: for example, an “H-type” electrolyzer using EDTA-Fe2+/Fe3+ as a redox couple (with 182-times lower membrane crossover than bare Fe2+/Fe3+) achieves stable co-production of CO (0.13 mmol·h−1·cm−2) and S at a current density of 8.5 mA·cm−2 [58]; a dual-flow electrolyzer with ultrathin interelectrode distance and CoPc-loaded gas diffusion electrodes (GDE) further boosts CO production rate by 7-fold and energy efficiency to 72.41%, while reducing the cost of natural gas purification by 0.03 $ per cubic meter [29,57]. Additionally, integrating solar energy via three-junction Si solar cells enables a solar-to-chemical (STC) conversion efficiency of 7.5–8.3%, and a novel CO2-assisted sulfur separation strategy avoids corrosive acids. CO2 adjusts electrolyte acidity to precipitate S8, with by-product NaHCO3, reused as CO2RR electrolyte, significantly enhancing atom utilization [13,58]. Practically, the system has been validated with simulated natural gas (Ar:CH4:CO2:H2S = 60:20:19:1 v/v), successfully converting concomitant CO2 and H2S into CO and S [58]. A two-electrode flow cell configuration even achieves 100 mA·cm−2 at only 1.5 V, far lower than conventional CO2RR-OER systems (~1.83 V), demonstrating great potential for industrial-scale carbon-neutral and resource-recycling applications [13,58].

6.2. SOR Coupled with Oxidizing Pollutants Treatment

Electrocatalytic SOR coupled with pollutant treatment (Figure 4), anodic oxidation for degrading sulfur-containing wastewater, and cathodic reduction in heavy metals and NO3 [87,94]. This system can precisely control the oxidation capacity and reaction pathways by adjusting parameters such as applied voltage and current density and is suitable for the simultaneous treatment of complex multi-component wastewater [94].
The coupling of reductive wastewater treatment in an electrochemical system refers to the oxidation and degradation of S2− at the anode and the reduction in heavy metals, NO3, or other oxidizing pollutants at the cathode during the operation of the electrochemical system. Specifically, the anode converts S2−/HS into S8 through direct or indirect oxidation pathways, while the cathode utilizes electrons generated from anodic SOR to drive the reduction in pollutants (such as nitrate/nitrite (NO3/NO2) to ammonia (NH3), hexavalent chromium (Cr6+) to low-toxic trivalent chromium (Cr3+), and vanadium ions (VO2+) to VO2+, realizing both pollutant detoxification and resource recovery [93,94,95]. This system is similar to the coupling of reductive wastewater treatment with the production of energy substances, but the difference is that heavy metals, NO3, or other oxidizing pollutants are used as electron acceptors at the cathode instead of reactions like HER or CO2RR [29]. Typically, anion or cation exchange membranes are employed to separate the anode and cathode compartments to prevent cross-reaction between oxidizing and reducing species while maintaining charge balance. Through this system, both reductive sulfur-containing wastewater and oxidizing heavy metal/nitrate-containing wastewater can be treated simultaneously, making it suitable for remediating complex multi-component industrial wastewater.

7. Conclusions and Perspectives

7.1. Conclusions

Hydrogen sulfide (H2S), a toxic by-product of fossil fuel processing and biomass energy development, poses severe threats to ecological stability, human health, and industrial infrastructure-yet its sulfur component represents a valuable resource with global annual demand exceeding 70 million tons. This review systematically synthesizes the latest progress in electrocatalytic H2S splitting for sulfur recovery, focusing on the sulfide oxidation reaction (SOR) as the core process, and reveals how electrochemical coupling systems have emerged as a transformative solution to address both H2S pollution and resource-energy nexus challenges.
From a mechanistic perspective, SOR pathways are clarified as two complementary routes: direct oxidation, where S2−/HS undergoes stepwise polymerization on anode surfaces to form S8 (with S* desorption as the rate-limiting step), and indirect oxidation, which leverages redox mediators (e.g., Fe3+/Fe2+, I/I3) to mitigate the primary bottleneck of electrode passivation caused by high-resistance S8 deposition. Accurate identification of SOR products-ranging from target S8 to intermediates (Sn2−) and by-products (S2O32−, SO42−)-has been enabled by advanced analytical techniques, including XRD, Raman spectroscopy, UV-vis spectrophotometry, and GC-MS; these methods not only validate product selectivity but also provide critical insights into reaction pathway dynamics, laying the groundwork for optimizing SOR conditions.
In terms of material design, SOR catalysts have evolved beyond traditional noble metals to include earth-abundant alternatives, such as transition metal sulfides (TMSs, e.g., NiS2 with “sulfur-repellent” surfaces), alloys (Pd-Co, high-entropy alloys), and metal-carbon composites (CoNi@NGs). The key design principle across these materials is modulating sulfur affinity: alloying and defect engineering tailor the d-band center to balance S2− adsorption and S8 desorption, while structural innovations (e.g., porous nanoarrays, heterojunctions) enhance electron transfer and mass diffusion. These advances have enabled catalysts to achieve high sulfur recovery rates (up to 92%) and long-term stability (over 100 h of operation), addressing critical performance limitations of early-stage materials.
Most notably, electrochemical coupling systems centered on SOR have expanded the technology’s value beyond sulfur recovery alone. By integrating SOR with cathode reactions such as hydrogen evolution reaction (HER), carbon dioxide reduction reaction (CO2RR), and oxygen reduction reaction (ORR), these systems simultaneously achieve “waste detoxification-sulfur recovery-high-value production”: for example, SOR-assisted HER reduces cell voltage to as low as 0.25 V (vs. traditional water electrolysis), while SOR-CO2RR co-produces CO (Faradaic efficiency > 83%) and S8, aligning with global “carbon peak and carbon neutrality” goals. For pollutant treatment, coupled systems also enable simultaneous remediation of sulfur-containing wastewater and reduction in oxidizing pollutants (e.g., NO3 to NH3, Cr6+ to Cr3+), demonstrating broad applicability in complex industrial scenarios.
Collectively, these advances highlight that electrocatalytic H2S splitting is no longer merely a pollution control technology but a multi-functional platform that bridges environmental protection, resource recycling, and energy production. Its ability to operate under mild conditions (ambient temperature/pressure) and adapt to low-concentration H2S (<5 vol%) addresses inherent limitations of traditional Claus processes, making it a promising candidate for industrial scaling.

7.2. Perspectives

Despite significant progress, electrocatalytic H2S splitting for sulfur recovery remains constrained by gaps between fundamental research and industrial application, and future efforts must focus on resolving core challenges while expanding the technology’s impact.
First, mechanistic understanding of SOR and passivation requires atomic-level resolution. Current studies have identified Sn2− as a key intermediate, but dynamic changes in adsorbed species (e.g., S3* and S3*) during reaction, as well as their interactions with catalyst surfaces, remain poorly characterized. Integrating in situ characterization techniques (e.g., in situ Raman/ATR-FTIR) with density functional theory (DFT) calculations will enable real-time tracking of intermediate evolution and quantification of energy barriers for rate-limiting steps. Furthermore, machine learning approaches can correlate catalyst electronic properties (d-band center, adsorption energy) with sulfur tolerance, establishing universal descriptors to accelerate the screening of anti-passivation catalysts—an advancement that would reduce the trial-and-error nature of current material design.
Second, catalyst innovation must prioritize low cost, sulfur resistance, and multifunctionality. While TMSs and high-entropy alloys show promise, their scalability is limited by synthesis complexity and reliance on rare elements (e.g., Pd in Pd-Co alloys). Future research should focus on earth-abundant elements (Fe, Co, Ni) and develop scalable synthesis methods (e.g., electrochemical deposition, hydrothermal synthesis) for high-performance catalysts. Structural design should further integrate “sulfur-repellent” surfaces (e.g., NiS2’s contact angle > 90°) with defect engineering to enhance both activity and stability; for example, doping Sn4+ into α-Fe2O3 introduces defect sites that inhibit deep oxidation of S2− to SO42−, a strategy that could be extended to other metal oxides. Additionally, developing bifunctional catalysts (e.g., TLNi@C HNSs for both SOR and HER) will simplify system design and reduce costs, a critical requirement for industrial adoption.
Third, electrochemical coupling systems need to advance toward integration and industrial adaptation. Current lab-scale systems often use batch reactors and ideal feedstocks (e.g., pure Na2S solutions), but industrial streams contain impurities (e.g., heavy metals in refinery wastewater) that degrade catalyst performance. Future systems should adopt membrane electrode assemblies (MEAs) and flow reactors to enhance mass transfer, suppress S8 deposition, and handle complex feedstocks; for instance, flow-type electrolyzers with anion exchange membranes have achieved industrial-level current densities (1 A cm−2) while maintaining high sulfur recovery. Redox mediator circulation (e.g., EDTA-Fe3+/Fe2+ with reduced membrane crossover) and CO2-assisted sulfur separation (using CO2 to adjust pH for S8 precipitation, with NaHCO3 reused as CO2RR electrolyte) can further improve atom efficiency and reduce secondary waste. Integrating solar energy (e.g., three-junction Si solar cells for photoelectrochemical systems) will also lower energy consumption, aligning the technology with renewable energy transitions.
Finally, the broader impact of this field lies in its potential to address interconnected global challenges: H2S pollution control, sulfur resource scarcity, and clean energy demand. By coupling SOR with waste-to-energy processes (e.g., microbial fuel cells for simultaneous electricity generation and sulfur recovery), the technology can contribute to circular economy models in the oil, gas, and coal chemical industries. Long-term vision should include cross-disciplinary collaboration-combining electrochemistry, materials science, and environmental engineering-to develop modular, scalable systems that can be deployed in both centralized (e.g., refineries) and decentralized (e.g., small-scale biogas plants) settings.
In summary, electrocatalytic H2S splitting for sulfur recovery stands at a critical juncture: with continued advances in mechanistic understanding, catalyst design, and system integration, it has the potential to become a cornerstone technology for sustainable pollution management and resource-energy synergy.

Author Contributions

C.C.: conceptualization, methodology, investigation, writing—original draft. X.G.: data curation, investigation. H.L.: investigation. Y.C.: conceptualization, project administration, funding acquisition, writing—review and editing. X.D.: formal analysis, visualization. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by Dazhou Key Laboratory of Advanced Technology for Fiber Materials (XWCL25ZA-02) Basalt Fiber and Composite Key Laboratory of Sichuan Province, Sichuan University of Arts and Science (XWFH-ZA-04), and Dazhou Ecological and Environmental Science Research Institute (2024HX027).

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Acknowledgments

This research was financially supported by the School of Chemistry and Chemical Engineering, Sichuan University of Arts and Science.

Conflicts of Interest

The authors confirm no conflict of interest.

References

  1. Qiao, L.; Bai, J.; Luo, T.; Li, J.H.; Zhang, Y.; Xia, L.G.; Zhou, T.S.; Xu, Q.J.; Zhou, B.X. High yield of H2O2 and efficient S recovery from toxic H2S splitting through a self-driven photoelectrocatalytic system with a microporous GDE cathode. Appl. Catal. B 2018, 238, 491–497. [Google Scholar] [CrossRef]
  2. Xiao, Z.H.; Lu, C.; Wang, J.; Qian, Y.Y.; Wang, B.W.; Zhang, Q.; Tang, A.D.; Yang, H.M. Bifunctional Co3S4 nanowires for robust sulfion oxidation and hydrogen generation with low power consumption. Adv. Funct. Mater. 2023, 33, 2212183. [Google Scholar] [CrossRef]
  3. Yu, Z.; Deng, Z.P.; Li, Y.; Wang, X.L. Advances in electrocatalyst design and mechanism for sulfide oxidation reaction in hydrogen sulfide splitting. Adv. Funct. Mater. 2024, 34, 2403435. [Google Scholar] [CrossRef]
  4. National Minerals Information Center. Sulfur Statistics and Information. Available online: https://www.usgs.gov/centers/national-minerals-information-center/sulfur-statistics-and-information (accessed on 29 October 2025).
  5. Wang, H.Q.; Song, S.; Zhang, Z.B.; Xin, L.Q.; Wang, T.Y.; Wang, L. A novel process of low-temperature fractionation combined with extractive distillation for H2S removal from natural gas. Sep. Purif. Technol. 2022, 302, 122102. [Google Scholar] [CrossRef]
  6. Song, Z.P.; Luo, T.F.; Ke, J.R.; Hu, S.S.; Yang, W.Q.; Ni, J.C.; Chen, X.P.; Chen, Z.H. A new-style pohotoelectrochemical sensing device based on NH2-UiO-66@Bi2O3 for the sensitive detection of hydrogen sulfide. Microchem. J. 2024, 206, 111669. [Google Scholar] [CrossRef]
  7. Sergienko, N.; Radjenovic, J. Manganese oxide-based porous electrodes for rapid and selective (electro)catalytic removal and recovery of sulfide from wastewater. Appl. Catal. B 2020, 267, 118608. [Google Scholar] [CrossRef]
  8. Pokorna, D.; Zabranska, J. Sulfur-oxidizing bacteria in environmental technology. Biotechnol. Adv. 2015, 33, 1246–1259. [Google Scholar] [CrossRef] [PubMed]
  9. Karlsen, V.B.; Dinamarca, C. Electrochemical and bioelectrochemical sulphide removal: A review. Rev. Environ. Sci. Biotechnol. 2024, 23, 989–1014. [Google Scholar] [CrossRef]
  10. Sergienko, N.; Lumbaque, E.C.; Radjenovic, J. (Electro)catalytic oxidation of sulfide and recovery of elemental sulfur from sulfide-laden streams. Water Res. 2023, 245, 120651. [Google Scholar] [CrossRef]
  11. Shao, X.H.; Huang, Y.X.; Wood, R.M.; Tarpeh, W.A. Electrochemical sulfate production from sulfide-containing wastewaters and integration with electrochemical nitrogen recovery. J. Hazard. Mater. 2024, 466, 133527. [Google Scholar] [CrossRef]
  12. Sun, X.M.; Huang, W.J.; Xu, H.M.; Qu, Z.; Yan, N.Q. Recovery of elemental sulfur direct from nonferrous flue gas by a low-temperature Claus process with H2S as the interim reducer. Fuel 2024, 359, 130352. [Google Scholar] [CrossRef]
  13. Teng, X.; Shi, K.; Chen, L.S.; Shi, J.L. Coupling electrochemical sulfion oxidation with CO2 reduction over highly dispersed p-Bi nanosheets and CO2-assisted sulfur extraction. Angew. Chem. 2024, 136, e202318585. [Google Scholar] [CrossRef]
  14. Zhou, Q.W.; Shen, Z.H.; Zhu, C.; Li, J.C.; Ding, Z.Y.; Wang, P.; Pan, F.; Zhang, Z.Y.; Ma, H.; Wang, S.Y.; et al. Nitrogen-doped CoP electrocatalysts for coupled hydrogen evolution and sulfur generation with low energy consumption. Adv. Mater. 2018, 30, 1800140. [Google Scholar] [CrossRef]
  15. Spedding, J.; Vujcich, M. Exchange of H2S between air and water. J. Geophys. Res. Ocean. 1982, 87, 8853–8856. [Google Scholar] [CrossRef]
  16. Yan, J.; Zhao, Z.X.; Wang, X.Y.; Xie, J. Hydrogen sulfide in seafood: Formation, hazards, and control. Trends Food Sci. Technol. 2024, 148, 104512. [Google Scholar] [CrossRef]
  17. Hedlund, F.H. Confined space hazards: Plain seawater, an insidious source of hydrogen sulfide. J. Occup. Environ. Hyg. 2023, 20, 322–328. [Google Scholar] [CrossRef]
  18. Lebrun, M.N.; Dorber, M.; Verones, F.; Henderson, A.D. Novel endpoint characterization factors for Life cycle impact assessment of terrestrial acidification. Ecol. Indic. 2025, 171, 113241. [Google Scholar] [CrossRef]
  19. Wang, X.; Zhang, R.; Yu, W. The effects of PM2.5 concentrations and relative humidity on atmospheric visibility in beijing. J. Geophys. Res. Atmos. 2019, 124, 2235–2259. [Google Scholar] [CrossRef]
  20. Ma, Q.X.; Zhang, C.Y.; Liu, C.; He, G.Z.; Zhang, P.; Li, H.; Chu, B.W.; He, H. A review on the heterogeneous oxidation of SO2 on solid atmospheric particles: Implications for sulfate formation in haze chemistry. Crit. Rev. Environ. Sci. Technol. 2023, 53, 1888–1911. [Google Scholar] [CrossRef]
  21. Fedorov, Y.A.; Mikhailenko, A.V.; Dotsenko, I.V. Sulfide sulfur in water objects with different mineralization. Water Resour. 2019, 46, S59–S64. [Google Scholar] [CrossRef]
  22. Abdul-Kadhim, R.M.; Oleiwi, M.S. Effect of agricultural sulfur and humic acid on some growth indicators of potato. IOP Conf. Ser. Earth Environ. Sci. 2025, 1487. [Google Scholar] [CrossRef]
  23. Bloem, E.; Haneklaus, S.; Kesselmeier, J.; Schnug, E. Sulfur fertilization and fungal infections affect the exchange of H2S and COS from agricultural crops. J. Agric. Food Chem. 2012, 60, 7588–7596. [Google Scholar] [CrossRef]
  24. Pang, H.Y.; Gao, T.; Zhao, W.K.; Liu, L.; Liu, E.Z.; Wen, H.Y.; Sun, T. Construction of MoSe2/NixSey/NF Schottky heterojunction as electrocatalyst for water splitting and electrochemical oxidation for sulfur recovery. Fuel 2024, 374, 132532. [Google Scholar] [CrossRef]
  25. PiÉPlu, A.; Saur, O.; Lavalley, J.C.; Legendre, O.; NÉDez, C. Claus catalysis and H2S selective oxidation. Catal. Rev. 1998, 40, 409–450. [Google Scholar] [CrossRef]
  26. Dutta, P.K.; Rabaey, K.; Yuan, Z.G.; Rozendal, R.A.; Keller, J. Electrochemical sulfide removal and recovery from paper mill anaerobic treatment effluent. Water Res. 2010, 44, 2563–2571. [Google Scholar] [CrossRef] [PubMed]
  27. Cho, W.B.; Kim, S.S.; Wie, J.J. Value-addition of wastes from petroleum refining process: Sulfur-rich polymers for sustainable and high-performanceoptical and energy applications. Acc. Mater. Res. 2024, 5, 625–639. [Google Scholar] [CrossRef]
  28. Zhang, J.; You, C.Y.; Lin, H.Z.; Wang, J. Electrochemical kinetic modulators in lithium–sulfur batteries: From defect-rich catalysts to single atomic catalysts. Energy Environ. Mater. 2022, 5, 731–750. [Google Scholar] [CrossRef]
  29. Yang, H.; Bai, J.; Zhou, T.S.; Zhou, C.H.; Xie, C.Y.; Zhang, Y.; Li, J.H.; Simchi, A.; Zhou, B.X. Electrochemical coupling conversion of sulfur-containing gaseous waste to treasure: A key review. Appl. Catal. A 2023, 654, 119085. [Google Scholar] [CrossRef]
  30. Chen, Y.; Liu, Y.; Gong, X.B.; Wang, J.L. Recovery of sulfur, generation of electricity and hydrogen peroxide from sulfion-rich wastewater using a novel self-driving photocatalytic fuel cell. Water Res. 2025, 275, 123232. [Google Scholar] [CrossRef]
  31. Huo, J.Y.; Jin, L.J.; Chen, C.C.; Chen, D.Y.; Xu, Z.C.; Wilfred, C.D.; Xu, Q.F.; Lu, J.M. Improving the sulfurophobicity of the NiS-doping CoS electrocatalyst boosts the low-energy-consumption sulfide oxidation reaction process. ACS Appl. Mater. Interfaces 2023, 15, 43976–43984. [Google Scholar] [CrossRef]
  32. Huo, J.Y.; Liu, Q.; Liu, X.F.; Cheng, X.F.; Chen, D.Y.; Li, N.J.; Liao, K.; Xu, Q.F.; Lu, J.M. Sulfur recovery assisted electrochemical water splitting for H2 production using CoMo-based nanorod arrays catalysts. ACS Mater. Lett. 2024, 6, 2633–2641. [Google Scholar] [CrossRef]
  33. Pei, Y.H.; Li, D.; Qiu, C.T.; Yan, L.; Li, Z.M.; Yu, Z.X.; Fang, W.Z.; Lu, Y.Y.; Zhang, B. High-entropy sulfide catalyst boosts energy-saving electrochemical sulfion upgrading to thiosulfate coupled with hydrogen production. Angew. Chem. Int. Ed. 2024, 63, e202411977. [Google Scholar] [CrossRef]
  34. Sergienko, N.; Irtem, E.; Gutierrez, O.; Radjenovic, J. Electrochemical removal of sulfide on porous carbon-based flow-through electrodes. J. Hazard. Mater. 2019, 375, 19–25. [Google Scholar] [CrossRef] [PubMed]
  35. Zhang, S.; Zhou, Q.W.; Shen, Z.H.; Jin, X.; Zhang, Y.C.; Shi, M.; Zhou, J.; Liu, J.G.; Lu, Z.D.; Zhou, Y.N.; et al. 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]
  36. Huang, C.; Yu, J.; Zhang, C.Y.; Cui, Z.B.; Chen, J.K.; Lai, W.H.; Lei, Y.J.; Nan, B.F.; Lu, X.; He, R.; et al. Electronic spin alignment within homologous NiS2/NiSe2 heterostructures to promote sulfur redox kinetics in lithium-sulfur batteries. Adv. Mater. 2024, 36, 2400810. [Google Scholar] [CrossRef]
  37. Xia, M.; Chen, R.; Zhu, X.; Liao, Q.; An, L.; Wang, Z.B.; He, X.F.; Jiao, L. A micro photocatalytic fuel cell with an air-breathing, membraneless and monolithic design. Sci. Bull. 2016, 61, 1699–1710. [Google Scholar] [CrossRef]
  38. Zong, X.; Chen, H.J.; Seger, B.; Pedersen, T.; Dargusch, M.S.; McFarland, E.W.; Li, C.; Wang, L.Z. Selective production of hydrogen peroxide and oxidation of hydrogen sulfide in an unbiased solar photoelectrochemical cell. Energy Environ. Sci. 2014, 7, 3347–3351. [Google Scholar] [CrossRef]
  39. Bai, J.; Zhang, B.; Li, J.H.; Zhou, B.X. Photoelectrocatalytic generation of H2 and S from toxic H2S by using a novel BiOI/WO3 nanoflake array photoanode. Front. Energy 2021, 15, 744–751. [Google Scholar] [CrossRef]
  40. Luo, T.; Bai, J.; Li, J.H.; Zeng, Q.Y.; Ji, Y.Z.; Qiao, L.; Li, X.Y.; Zhou, B.X. Self-driven photoelectrochemical splitting of H2S for S and H2 recovery and simultaneous electricity generation. Environ. Sci. Technol. 2017, 51, 12965–12971. [Google Scholar] [CrossRef] [PubMed]
  41. Zong, X.; Han, J.F.; Seger, B.; Chen, H.J.; Lu, G.Q.; Li, C.; Wang, L.Z. An integrated photoelectrochemical-chemical loop for solar-driven overall splitting of hydrogen sulfide. Angew. Chem. Int. Ed. 2014, 53, 4399–4403. [Google Scholar] [CrossRef] [PubMed]
  42. Liu, Y.; Wang, J.L. Multivalent metal catalysts in Fenton/Fenton-like oxidation system: A critical review. Chem. Eng. J. 2023, 466, 143147. [Google Scholar] [CrossRef]
  43. Zheng, D.; Yang, X.Q.; Qu, D.Y. Reaction between lithium anode and polysulfide ions in a lithium-sulfur battery. ChemSusChem 2016, 9, 2348–2350. [Google Scholar] [CrossRef]
  44. He, D.T.; Yang, P.J.; Yang, K.Z.; Qiu, J.S.; Wang, Z.Y. Long-lasting hybrid seawater electrolysis enabled by anodic mass transport intensification for energy-saving hydrogen production. Adv. Funct. Mater. 2024, 34, 2407601. [Google Scholar] [CrossRef]
  45. Avetisyan, K.; Zweig, I.; Luther, G.W.; Kamyshny, A. Kinetics and mechanism of polysulfides and elemental sulfur formation by a reaction between hydrogen sulfide and δ-MnO2. Geochim. Cosmochim. Acta 2021, 313, 21–37. [Google Scholar] [CrossRef]
  46. Liu, T.; Liu, Y.; Wang, J.W.; Wang, J.L. Photo-generated hydrated electrons for selective reduction of nitrate to dinitrogen: Selectivity and mechanism. Chem. Eng. J. 2024, 497, 154860. [Google Scholar] [CrossRef]
  47. Zhang, M.; Guan, J.; Tu, Y.C.; Chen, S.M.; Wang, Y.; Wang, S.H.; Yu, L.; Ma, C.; Deng, D.H.; Bao, X.H. Highly efficient H2 production from H2S via a robust graphene-encapsulated metal catalyst. Energy Environ. Sci. 2020, 13, 119–126. [Google Scholar] [CrossRef]
  48. Huang, Q.Q.; Chang, Q.Y.; Luo, Z.Y.; Xie, M.F.; Jiang, N.B.; Zhang, X.N.; Zhou, M.; Zhang, Y.H.; Xiao, P. Superimposing effect of electrocatalytic activity and photocatalytic activity for amorphous NiFe-sulfides photoanode in sulfion oxidation reaction. Surf. Interfaces 2024, 54, 105163. [Google Scholar] [CrossRef]
  49. Kim, K.Y.; Han, J.I. Carbon supported bimetallic Pd–Co catalysts for alkaline sulfide oxidation in direct alkaline sulfide fuel cell. Int. J. Hydrogen Energy 2015, 40, 4567–4572. [Google Scholar] [CrossRef]
  50. Haner, J.; Bejan, D.; Bunce, N.J. Electrochemical oxidation of sulfide ion at a Ti/IrO2-Ta2O5 anode in the presence and absence of naphthenic acids. J. Appl. Electrochem. 2009, 39, 1733–1738. [Google Scholar] [CrossRef]
  51. Pikaar, I.; Rozendal, R.A.; Yuan, Z.G.; Keller, J.; Rabaey, K. Electrochemical sulfide oxidation from domestic wastewater using mixed metal-coated titanium electrodes. Water Res. 2011, 45, 5381–5388. [Google Scholar] [CrossRef]
  52. Pei, Y.H.; Cheng, J.; Zhong, H.; Pi, Z.F.; Zhao, Y.; Jin, F.M. 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]
  53. Bedoya Lora, F.; Hankin, A.; Kelsall, G.H. Photo-electrochemical hydrogen sulfide splitting using SnIV-doped hematite photo-anodes. Electrochem. Commun. 2016, 68, 19–22. [Google Scholar] [CrossRef]
  54. Bessekhouad, Y.; Mohammedi, M.; Trari, M. Hydrogen photoproduction from hydrogen sulfide on Bi2S3 catalyst. Sol. Energy Mater. Sol. Cells 2002, 73, 339–350. [Google Scholar] [CrossRef]
  55. Ren, J.T.; Chen, L.; Wang, H.Y.; Tian, W.W.; Wang, L.; Sun, M.L.; Feng, Y.; Zhai, S.X.; Yuan, Z.Y. Self-powered hydrogen production with improved energy efficiency via polysulfides redox. ACS Nano 2023, 17, 25707–25720. [Google Scholar] [CrossRef] [PubMed]
  56. Li, F.L.; Ng, W.K.; Qiao, W.; Yang, Y.T.; Li, X.G.; Xi, B.J.; Yu, Y.; Fang, J.Y.; Li, P.; Xiong, S.L. Intercalation chemistry awakens transition metal hydroxide for boosted and sustained electrocatalytic sulfion oxidation. Angew. Chem. 2025, 137, e202511402. [Google Scholar] [CrossRef]
  57. Zhang, B.; Bai, J.; Zhang, Y.; Zhou, C.; Wang, P.B.; Zha, L.N.; Li, J.H.; Simchi, A.; Zhou, B.X. High yield of CO and synchronous S recovery from the conversion of CO2 and H2S in natural gas based on a novel electrochemical reactor. Environ. Sci. Technol. 2021, 55, 14854–14862. [Google Scholar] [CrossRef]
  58. Ma, W.G.; Wang, H.; Yu, W.; Wang, X.M.; Xu, Z.Q.; Zong, X.; Li, C. Achieving simultaneous CO2 and H2S conversion via a coupled solar-driven electrochemical approach on non-precious-metal catalysts. Angew. Chem. Int. Ed. 2018, 57, 3473–3477. [Google Scholar] [CrossRef]
  59. Liu, X.D.; Wang, X.Y.; Long, J.Q.; Xie, X.F.; Wu, L.; Wang, Z.J.; Fu, Y.X.; Chen, H.; Xiang, K.S.; Liu, H. Research on the selective electrocatalytic reduction of SO2 to recover S0 by Pb electrode. Metals 2023, 13, 569. [Google Scholar] [CrossRef]
  60. Jiao, S.L.; Fu, X.W.; Huang, H.W. Descriptors for the evaluation of electrocatalytic reactions: D-band theory and beyond. Adv. Funct. Mater. 2022, 32, 2107651. [Google Scholar] [CrossRef]
  61. Hua, W.X.; Shang, T.X.; Li, H.; Sun, Y.F.; Guo, Y.; Xia, J.Y.; Geng, C.N.; Hu, Z.H.; Peng, L.K.; Han, Z.Y.; et al. Optimizing the p charge of S in p-block metal sulfides for sulfur reduction electrocatalysis. Nat. Catal. 2023, 6, 174–184. [Google Scholar] [CrossRef]
  62. Sun, Y.F.; Wang, J.Y.; Shang, T.X.; Li, Z.J.; Li, K.H.; Wang, X.W.; Luo, H.R.; Lv, W.; Jiang, L.L.; Wan, Y. Counting d-orbital vacancies of transition-metal catalysts for the sulfur reduction reaction. Angew. Chem. Int. Ed. 2023, 62, e202306791. [Google Scholar] [CrossRef]
  63. He, R.Z.; Huang, X.Y.; Feng, L.G. Recent progress in transition-metal sulfide catalyst Regulation for improved oxygen evolution Reaction. Energy Fuels 2022, 36, 6675–6694. [Google Scholar] [CrossRef]
  64. Xia, H.; Shi, Z.D.; Gong, C.S.; He, Y.M. Recent strategies for activating the basal planes of transition metal dichalcogenides towards hydrogen production. J. Mater. Chem. A 2022, 10, 19067–19089. [Google Scholar] [CrossRef]
  65. Li, Y.F.; Duan, Y.X.; Zhang, K.; Yu, W.Z. 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]
  66. Luo, M.; Yang, J.T.; Li, X.G.; Eguchi, M.; Yamauchi, Y.; Wang, Z.L. Insights into alloy/oxide or hydroxide interfaces in Ni–Mo-based electrocatalysts for hydrogen evolution under alkaline conditions. Chem. Sci. 2023, 14, 3400–3414. [Google Scholar] [CrossRef]
  67. Wang, Y.; Mi, J.X.; Wu, Z.S. Recent status and challenging perspective of high entropy oxides for chemical catalysis. Chem Catal. 2022, 2, 1624–1656. [Google Scholar] [CrossRef]
  68. Lei, Y.; Zhang, L.; Xu, W.; Xiong, C.; Chen, W.; Xiang, X.; Zhang, B.; Shang, H. Carbon-supported high-entropy Co-Zn-Cd-Cu-Mn sulfide nanoarrays promise high-performance overall water splitting. Nano Res. 2022, 15, 6054–6061. [Google Scholar] [CrossRef]
  69. Buckingham, M.A.; Ward-O’Brien, B.; Xiao, W.C.; Li, Y.; Qu, J.; Lewis, D.J. High entropy metal chalcogenides: Synthesis, properties, applications and future directions. Chem. Commun. 2022, 58, 8025–8037. [Google Scholar] [CrossRef]
  70. Wang, Q.Q.; Li, J.Q.; Li, Y.J.; Shao, G.M.; Jia, Z.; Shen, B.L. Non-noble metal-based amorphous high-entropy oxides as efficient and reliable electrocatalysts for oxygen evolution reaction. Nano Res. 2022, 15, 8751–8759. [Google Scholar] [CrossRef]
  71. Guan, S.X.; Liang, H.; Wang, Q.M.; Tan, L.J.; Peng, F. Synthesis and phase stability of the high-entropy carbide (Ti0.2Zr0.2Nb0.2Ta0.2Mo0.2)C under extreme conditions. Inorg. Chem. Front. 2021, 60, 3807–3813. [Google Scholar] [CrossRef]
  72. Kim, K.; Han, J.-I. Carbon-supported bimetallic Pd-Ir catalysts for alkaline sulfide oxidation in direct alkaline sulfide fuel cell. J. Appl. Electrochem. 2015, 45, 533–539. [Google Scholar] [CrossRef]
  73. Li, J.C.; Hou, P.X.; Zhao, S.Y.; Liu, C.; Tang, D.M.; Cheng, M.; Zhang, F.; Cheng, H.M. A 3D bi-functional porous N-doped carbon microtube sponge electrocatalyst for oxygen reduction and oxygen evolution reactions. Energy Environ. Sci. 2016, 9, 3079–3084. [Google Scholar] [CrossRef]
  74. Deng, Z.P.; Gong, M.X.; Gong, Z.; Wang, X.L. Mesoscale mass transport enhancement on well-defined porous carbon platform for electrochemical H2O2 synthesis. Nano Lett. 2022, 22, 9551–9558. [Google Scholar] [CrossRef]
  75. Singh, R.; Chaudhary, S.; Yadav, S.; Patil, S.A. Bioelectrocatalytic sulfide oxidation by a haloalkaliphilic electroactive microbial community dominated by Desulfobulbaceae. Electrochim. Acta 2022, 423, 140576. [Google Scholar] [CrossRef]
  76. Słowiński, D.; Świerczyńska, M.; Grzelakowska, A.; Szala, M.; Kolińska, J.; Romański, J.; Podsiadły, R. Hymecromone naphthoquinone ethers as probes for hydrogen sulfide detection. Dye. Pigm. 2021, 196, 109765. [Google Scholar] [CrossRef]
  77. Xu, H.; Yan, W. Studies on the preparation of high efficient Ti/PbO2 electrode and degradation of acid red G. China Environ. Sci. 2017, 37, 2591–2598. [Google Scholar]
  78. Qi, F.L.; Wang, X.M.; Zhu, P.P.; Li, Q.D.; Wang, S.K.; Zhao, J.J. Asymmetric Fe2+/Fe3+-mediated flow-electrode capacitive deionization for the removal of chloride ions in reclaimed water. ACS Sustain. Chem. Eng. 2024, 12, 8609–8619. [Google Scholar] [CrossRef]
  79. Xu, N.; Wang, T.L.; Li, W.J.; Wang, Y.; Chen, J.J.; Liu, J. Tuning redox potential of anthraquinone-2-sulfonate (AQS) by chemical modification to facilitate electron transfer from electrodes in shewanella oneidensis. Front. Bioeng. Biotechnol. 2021, 9, 705414. [Google Scholar] [CrossRef]
  80. Blázquez, E.; Gabriel, D.; Baeza, J.A.; Guisasola, A.; Freguia, S.; Ledezma, P. Recovery of elemental sulfur with a novel integrated bioelectrochemical system with an electrochemical cell. Sci. Total Environ. 2019, 677, 175–183. [Google Scholar] [CrossRef]
  81. Rabaey, K.; Van de Sompel, K.; Maignien, L.; Boon, N.; Aelterman, P.; Clauwaert, P.; De Schamphelaire, L.; Pham, H.T.; Vermeulen, J.; Verhaege, M.; et al. Microbial Fuel Cells for Sulfide Removal. Environ. Sci. Technol. 2006, 40, 5218–5224. [Google Scholar] [CrossRef]
  82. Cao, B.C.; Zhao, Z.P.; Peng, L.L.; Shiu, H.Y.; Ding, M.N.; Song, F.; Guan, X.; Lee, C.K.; Huang, J.; Zhu, D.; et al. Silver nanoparticles boost charge-extraction efficiency in Shewanella microbial fuel cells. Science 2021, 373, 1336–1340. [Google Scholar] [CrossRef]
  83. Wu, J.; Cao, J.p.; Bi, H.l.; Zhang, J.; Cao, Q. Liquid-solid contact electrification and its effect on the formation of electric double layer: An atomic-level investigation. Nano Energy 2023, 111, 108442. [Google Scholar] [CrossRef]
  84. Wang, J.; Liu, X.M.; Ma, C.B.; Duan, X.G.; Li, S.; Li, N.; Liu, W.; Li, Y.; Fan, X.B.; Peng, W.C. Asymmetric S heteroatom coordinated dual-atom catalysts and coupled anodic sulfion oxidation to boost electrocatalysis oxygen reduction. Adv. Funct. Mater. 2025, 35, 2420157. [Google Scholar] [CrossRef]
  85. Mishra, P.; Saravanan, P.; Packirisamy, G.; Jang, M.; Wang, C.Y. A subtle review on the challenges of photocatalytic fuel cell for sustainable power production. Int. J. Hydrogen Energy 2021, 46, 22877–22906. [Google Scholar] [CrossRef]
  86. Wang, W.X.; Mao, Q.Q.; Deng, K.; Yu, H.J.; Wang, Z.Q.; Xu, Y.; Li, X.N.; Wang, L.; Wang, H.J. Sulfur-induced low crystallization of ultrathin Pd nanosheet arrays for sulfur Ion degradation-assisted energy-efficient H2 production. Small 2023, 19, 2207852. [Google Scholar] [CrossRef]
  87. Wu, S.; Zhang, Y.; Xie, Y.J.; Bai, X.; Fan, G.Y.; Yu, X.J. Ruthenium nanocluster-regulated electronic behaviors of nickel hydroxides for boosted coupled-upgrading of nitrite and sulfide pollutants while outputting energy through zinc-nitrite battery. J. Alloys Compd. 2025, 1039, 183232. [Google Scholar] [CrossRef]
  88. Zhang, Y.; Zhu, Y.Y.; Zhao, D.L.; Fan, G.Y.; Long, Y. Synergy electrocatalysis between palladium and cobalt aligned on carbon nanosheet architectures for nitrate-to-ammonia conversion coupled with sulfur ion oxidation. J. Colloid Interface Sci. 2026, 701, 138688. [Google Scholar] [CrossRef]
  89. Baek, K.; Li, T.L.; Lee, C.; Li, W.Z.; Kim, K. MOF-derived CoSx as a bifunctional electrocatalyst for efficient sulfide oxidation and coupled ammonia synthesis. Green Chem. 2025, 27, 8637–8648. [Google Scholar] [CrossRef]
  90. Yang, M.; Hou, X.; Fu, R.; Wang, Z.; Hu, C.; Hu, G.; Zhuo, L.; Liu, X. Cation doping driven performance optimization of MoS2 nanoarrays for nitrate and sulfide co-electrolysis. Nano Res. 2025, 18, 94907778. [Google Scholar] [CrossRef]
  91. Yang, M.S.; Wei, T.R.; Zeng, C.H.; Zhang, J.W.; Liu, Y.F.; Luo, J.; Hu, G.Z.; Liu, X.J. CoNiOOH nanosheets array enables highly effective value-added chemicals production via nitrite and sulfide electrolysis. Chem. Eng. J. 2024, 498, 155799. [Google Scholar] [CrossRef]
  92. Zhang, Z.N.; Wang, X.H.; Du, Q.Y.; Hong, Q.L.; Ai, X.; Chen, Y.; Li, S.N. Defect-rich AuCu/CuS nanowires heterojunction for light-enhanced sulfur Ion electrooxidation coupled nitrite electroreduction. Adv. Energy Mater. 2025, 15, 2500176. [Google Scholar] [CrossRef]
  93. Wang, X.H.; Hong, Q.L.; Shao, L.Y.; Zhai, Q.G.; Jiang, Y.C.; Ai, X.; Chen, Y.; Li, S.N. Copper-nickel oxide nanosheets with atomic thickness for high-efficiency sulfur Ion electrooxidation assisted nitrate electroreduction to ammonia. Adv. Funct. Mater. 2024, 34, 2408834. [Google Scholar] [CrossRef]
  94. Kijjanapanich, P.; Kijjanapanich, P.; Annachhatre, A.P.; Esposito, G.; Lens, P.N.L. Spontaneous electrochemical treatment for sulfur recovery by a sulfide oxidation/vanadium(V) reduction galvanic cell. J. Environ. Manag. 2015, 149, 263–270. [Google Scholar] [CrossRef] [PubMed]
  95. Saad, E.G.; Zewail, T.M.; Zatout, A.A.; El-Ashtoukhy, E.S.Z.; Abdel-Aziz, M.H. Electrochemical removal of sulfide ions and recovery of sulfur from sulfide ions containing wastes. J. Ind. Eng. Chem. 2021, 94, 390–396. [Google Scholar] [CrossRef]
  96. Xiong, Y.R.; Wang, L.L.; Ning, P.; Luo, J.F.; Li, X.; Yuan, L.; Xie, Y.B.; Ma, Y.X.; Wang, X.Q. Constructing oxygen vacancy-enriched Fe3O4@MnO2 core-shell nanoplates for highly efficient catalytic oxidation of H2S in blast furnace gas. Sep. Purif. Technol. 2024, 336, 126234. [Google Scholar] [CrossRef]
  97. Peng, W.; Tan, H.T.; Liu, X.T.; 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]
  98. Wu, Q.Y.; Cao, J.J.; Wang, X.; Liu, Y.; Zhao, Y.J.; Wang, H.; Liu, Y.; Huang, H.; Liao, F.; Shao, M.W.; et al. A metal-free photocatalyst for highly efficient hydrogen peroxide photoproduction in real seawater. Nat. Commun. 2021, 12, 483. [Google Scholar] [CrossRef] [PubMed]
  99. Chen, Y.; Liu, Y.; Gong, X.B.; Wang, J.L. Photocatalytic degradation of chlorinated organic pollutants by ZnS@ZIF-8 composite through hydrogen peroxide generation by activating dioxygen under simulated sunlight irradiation. J. Colloid Interface Sci. 2024, 654, 1417–1430. [Google Scholar] [CrossRef]
  100. Liu, L.j.; Gao, M.Y.; Yang, H.f.; Wang, X.y.; Li, X.b.; Cooper, A.I. Linear conjugated polymers for solar-driven hydrogen peroxide production: The importance of catalyst stability. J. Am. Chem. Soc. 2021, 143, 19287–19293. [Google Scholar] [CrossRef]
  101. Tan, N.; Yang, Z.; Gong, X.B.; Wang, Z.R.; Fu, T.; Liu, Y. In situ generation of H2O2 using MWCNT-Al/O2 system and possible application for glyphosate degradation. Sci. Total Environ. 2019, 650, 2567–2576. [Google Scholar] [CrossRef]
  102. Gong, X.B.; Yang, Z.; Peng, L.; Zhou, A.L.; Liu, Y.L.; Liu, Y. In-situ synthesis of hydrogen peroxide in a novel Zn-CNTs-O2 system. J. Power Sources 2018, 378, 190–197. [Google Scholar] [CrossRef]
  103. Bo, X.Y.; Yang, C.J.; Li, B.L.; Fan, Y.K.; Li, J.Y.; Huang, H.M.; Kou, J.H.; Lu, C.H. Reverse quantum wells in gradient-doped CdS for photocatalytic H2O2 production. Inorg. Chem. Front. 2025, 12, 5177–5188. [Google Scholar] [CrossRef]
  104. Zhang, Y.Y.; Zheng, X.; Su, H.; Ling, Y.; Guo, R.; Zhang, M.S.; Wang, Q.X.; Niu, L. Regulating the bubble-water/catalyst interface microenvironment for accelerated electrosynthesis of H2O2 via optimizing oxygen functional groups on carbon black. Green Chem. 2025, 27, 3315–3325. [Google Scholar] [CrossRef]
  105. Garg, K.; Kumar, M.; Kaur, S.; Nagaiah, T.C. Electrochemical production of hydrogen from hydrogen sulfide using cobalt cadmium sulfide. ACS Appl. Mater. Interfaces 2023, 15, 27845–27852. [Google Scholar] [CrossRef] [PubMed]
  106. Zhu, Y.; Wang, S.; Chen, Y.X.; Zhang, Y.Y.; Feng, Y.F.; Zhang, G.Q. Screened d-p orbital hybridization in turing structure of confined nickel for sulfion oxidation accelerated hydrogen production. Angew. Chem. Int. Ed. 2025, 64, e202419572. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Mechanistic diagram of SOR for recovery S8.
Figure 1. Mechanistic diagram of SOR for recovery S8.
Catalysts 15 01019 g001
Figure 2. The analysis methods of SOR products. (a) XRD pattern [30], (b) Raman spectra [2], (c) UV-vis spectra [13], (d) GC-MS [43].
Figure 2. The analysis methods of SOR products. (a) XRD pattern [30], (b) Raman spectra [2], (c) UV-vis spectra [13], (d) GC-MS [43].
Catalysts 15 01019 g002
Figure 3. SOR coupled with the production of energy substances. (a) SOR coupled with ORR [38]. (b) SOR coupled with HER [86]. (c,d) SOR coupled with CO2RR [13,58].
Figure 3. SOR coupled with the production of energy substances. (a) SOR coupled with ORR [38]. (b) SOR coupled with HER [86]. (c,d) SOR coupled with CO2RR [13,58].
Catalysts 15 01019 g003
Figure 4. SOR coupled with oxidizing pollutant treatments. (a,b) SOR coupled with NO3RR [87,90], (c) SOR coupled with reduced VO2+ [94], and (d) SOR coupled with reduced Cr3+ [95].
Figure 4. SOR coupled with oxidizing pollutant treatments. (a,b) SOR coupled with NO3RR [87,90], (c) SOR coupled with reduced VO2+ [94], and (d) SOR coupled with reduced Cr3+ [95].
Catalysts 15 01019 g004
Table 1. Different types of electrodes for SOR.
Table 1. Different types of electrodes for SOR.
Electrodes TypeAnodesActive Species of SORProducts of SORReferences
Direct oxidation electrodesCoNi@NGsCo3+/Co2+, Ni3+/Ni2+S8[47]
NiFeS/ZnIn2S4Ni3+/Ni2+S8[48]
Pd8Co2/CPd2+/Pd0, Co3+/Co2+, OHadsS8[49]
Ti/IrO2-Ta2O5Ir4+/Ir3+S8[50]
MMO/TiO2, OHS0, S2O32−, SO42−, SO32−[51]
Ir/Ta/TiIr4+/Ir3+, OH/O2S0, SO42−, S2O32−[11]
Controllable Sn2− producing electrodesCuCoNiMnCrSx/NFCu2+/Cu+, Mn3+/Cr3+Sn2−[33]
Cu2S/NFCu2+/Cu+Sn2−[52]
Sn4+-doped α-Fe2O3Fe3+/Fe2+Sn2−[53]
Bi2S3Bi3+/Bi2+Sn2−[54]
HEA-Mo2C/HPCSx2−/Sx+12−Sn2−[55]
NF/Mo-Co(OH)2Co(OH)2Sn2−[56]
Medium circulation-driven indirect oxidation electrodesGraphite feltI/I3S8[57]
WO3/SiPVCI/I3S8[1]
WO3/FTOI/I3S0[40]
n-Si@PEDOT/p-Si@PtFe3+/Fe2+S0[41]
G/GCSEDTA-Fe3+/Fe2+S0[58]
Table 2. Different electrochemical systems for SOR.
Table 2. Different electrochemical systems for SOR.
Electrochemical Coupling SystemsAnodesProducts of SORCathodesProducts in CathodesReferences
SOR coupled with the production of energy substancesSn4+-doped α-Fe2O3Sn2−PtH2[53]
Bi2S3Sn2−PtH2[54]
WO3/SiPVCSn2−GDEH2O2[1]
p-SiSTiO2/Ti/n+p-SiH2O2[38]
WO3/FTOS0Pt/SiPVCH2[40]
Cu2S/NFSn2−Pt/GrH2[52]
HEA-Mo2C/HPCSn2−HEA-Mo2C/HPCH2[55]
CoS-NFSn2−CoGa-NS-CH2O2[84]
CoNiS2/NFS0/S2O32−PtH2[85]
a/c S-Pd NSA/NFSn2−a/c S-Pd NSA/NFH2[86]
NiS2S0NiS2H2[35]
Co-S NSsSn2−p-Bi NSsHCOOH[13]
n-Si@PEDOT/p-Si@PtS0p-Si@PtH2[41]
NiS–CoS/CNFS0PtH2[31]
G/GCSS0ZnO@GrCO[58]
SOR coupled with oxidizing pollutants treatmentRu-Ni(OH)2/NFS0Ru-Ni(OH)2/NFNH3[87]
Pd-Co@NC/CCS0Pd–Co@NC/CCNH3[88]
MOF-derived CoSxS0MOF-derived CoSxNH3[89]
Ni-MoS2@ACFS0Ni-MoS2@ACFNH3[90]
CoNiOOH/CCS0CoNiOOH/CCNH3[91]
AuCu/CuS NWsS0AuCu/CuS NWsNH3[92]
Cu-NiO UTNSsS0Cu–NiO UTNSsNH3[93]
GrS0GrVO2+[94]
GrS0GrCr3+[95]
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

Chen, C.; Geng, X.; Liu, H.; Chen, Y.; Deng, X. Advances in Electrocatalytic Hydrogen Sulfide Splitting for Sulfur Recovery: From Reaction Mechanisms to Application. Catalysts 2025, 15, 1019. https://doi.org/10.3390/catal15111019

AMA Style

Chen C, Geng X, Liu H, Chen Y, Deng X. Advances in Electrocatalytic Hydrogen Sulfide Splitting for Sulfur Recovery: From Reaction Mechanisms to Application. Catalysts. 2025; 15(11):1019. https://doi.org/10.3390/catal15111019

Chicago/Turabian Style

Chen, Chuntan, Xiangyong Geng, Hepei Liu, Yong Chen, and Xinshuang Deng. 2025. "Advances in Electrocatalytic Hydrogen Sulfide Splitting for Sulfur Recovery: From Reaction Mechanisms to Application" Catalysts 15, no. 11: 1019. https://doi.org/10.3390/catal15111019

APA Style

Chen, C., Geng, X., Liu, H., Chen, Y., & Deng, X. (2025). Advances in Electrocatalytic Hydrogen Sulfide Splitting for Sulfur Recovery: From Reaction Mechanisms to Application. Catalysts, 15(11), 1019. https://doi.org/10.3390/catal15111019

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