Boosting the efficiency of electrocatalytic water splitting via in situ grown transition metal sulfides: a review

Haowei Jia a, Linghui Meng a, Yile Lu a, Tianyue Liang a, Yu Yuan a, Yifan Hu a, Zekun Dong a, Yingze Zhou *a, Peiyuan Guan *a, Lu Zhou a, Chao Liu a, Mengyao Li a, Tao Wan a, Bing-Jie Ni b, Zhaojun Han *c and Dewei Chu a
aSchool of Materials Science and Engineering, The University of New South Wales, Sydney 2052, Australia. E-mail: yingze.zhou@unsw.edu.au; peiyuan.guan@unsw.edu.au
bCivil and Environmental Engineering, The University of New South Wales, Sydney 2052, Australia
cSchool of Mechanical, Medical and Process Engineering, Queensland University of Technology, Brisbane 4000, Australia. E-mail: zhaojun.han@qut.edu.au

Received 1st September 2024 , Accepted 26th September 2024

First published on 27th September 2024


Abstract

The growing needs for sustainable and efficient energy sources have heightened interest in electrocatalytic water splitting (EWS), a promising method for hydrogen production as a clean and renewable energy carrier. EWS, which splits water molecules into hydrogen and oxygen, faces efficiency challenges due to the slow kinetics of the oxygen evolution reaction (OER) at the anode and the hydrogen evolution reaction (HER) at the cathode. Developing effective electrocatalysts is essential to overcome these limitations. Among the various electrocatalysts studied, transition metal sulfides (TMSs) have garnered significant attention due to their low cost and abundant active sites. Despite the superior electrochemical performance of TMSs compared to other materials, their inherent low conductivity and sluggish reaction kinetics remain major challenges. Recent advancements have focused on the in situ growth of TMSs on conductive substrates to enhance electron transfer and overall catalytic performance, eliminating the need for polymer binders and improving electrode stability. This review provides an in-depth analysis of the key aspects involved in the synthesis of in situ grown TMS electrodes, including the selection of TMS active materials, various substrates, and preparation strategies. The review then offers a comprehensive overview of different types of in situ grown TMS electrodes, with a focus on the most extensively researched materials: molybdenum sulfides, cobalt sulfides, nickel sulfides, and their composites. Finally, the limitations and future perspectives are discussed, highlighting potential directions for advancing the development of in situ grown TMS catalysts.


1. Introduction

The pursuit of sustainable and efficient energy sources has significantly intensified the focus on electrocatalytic water splitting (EWS), a promising process for hydrogen production as a clean and renewable energy carrier. EWS entails the splitting of water molecules into hydrogen and oxygen by applying electrical energy. However, its efficiency is limited by the slow kinetics of the oxygen evolution reaction (OER) at the anode and the hydrogen evolution reaction (HER) at the cathode (Fig. 1a). Therefore, the development of effective electrocatalysts is crucial for improving the efficiency of EWS.
image file: d4ta06197g-f1.tif
Fig. 1 (a) Schematic representation of the mechanisms of the HER and OER in both alkaline and acidic media; (b) schematic diagram comparing in situ grown and ex situ grown electrodes; (c) trends in scientific publications by category from 2014 to 2023.

To date, a wide range of electrocatalysts have been extensively studied to improve the EWS reaction efficiency, including noble metal-based electrodes such as platinum (Pt), iridium dioxide (IrO2), ruthenium dioxide (RuO2),1,2 transition metal oxides (TMOs) such as molybdenum trioxide (MoO3), cobalt oxide (CoO), manganese dioxide (MnO2)3–6 and TMSs such as molybdenum disulfide (MoS2), vanadium sulfides (VS2), nickel sulfides (Ni3S2).7–9 Among these, TMSs have garnered particular attention due to their low cost and abundant active reaction sites.10 While noble metal-based materials are well-known for their high catalytic efficiency, their high cost and non-renewable nature considerably limit their broad practical application. Additionally, compared to TMOs, TMSs often exhibit superior electrochemical performance, which is mainly attributed to the following reasons: (1) oxygen is more electronegative than sulfur, meaning that when oxygen atoms are replaced by sulfur in TMSs, the bonding becomes less polar. Sulfur has a larger atomic radius and a lower electronegativity, which weakens the metal–sulfur bonds compared to metal–oxygen bonds. (2) The replacement of oxygen with sulfur tends to narrow the bandgap of the material. A narrower bandgap improves electrical conductivity, which is crucial for catalytic applications. (3) Sulfur atoms provide different surface electronic states compared to oxygen, which can create more favorable conditions for adsorption of reactants. This can increase the density and reactivity of active sites, leading to enhanced catalytic activity in various processes.11–14 These features make TMS materials particularly well-suited for applications in electrocatalysis (Table 1).15,16

Table 1 The conclusion summarizing the advantages and disadvantages of noble metal-based electrodes, TMO electrodes and TMS electrodes
Advantages Disadvantages
Noble metal-based electrodes • Outstanding conductivity • High cost
• Superior catalytic activity • Non-renewable
• Excellent stability
TMO electrodes • Good conductivity • Low catalytic activity
• Low cost • Unstable structure
• Simple preparation methods
TMS electrodes • Low cost • Low conductivity
• Good catalytic activity • Sluggish reaction kinetics
• Adjustable structures


However, TMS materials still face certain drawbacks, including inherent low conductivity and sluggish reaction kinetics.17–19 To address these challenges and enhance the electrocatalytic efficiency of TMSs, conductive materials are often integrated during electrode preparation to improve the electrochemical performance of TMSs.20,21 Nevertheless, the use of inert polymer binders can obstruct the active sites of TMS electrocatalysts, significantly diminishing their catalytic activity during EWS. In response to this issue, the in situ growth of TMSs on conductive substrates has attracted considerable attention. This method ensures close contact between the active materials and the substrate, thereby improving electron transfer and enhancing overall catalytic performance.22,23

Recent research has increasingly focused on in situ grown electrodes due to several key advantages: Firstly, in situ grown electrodes eliminate the need for polymer binders and slurry ink, thereby reducing the cost of electrode preparation.24 Secondly, improved contact between the active materials and substrates results in superior electron transfer and mass transport, which can facilitate the reaction kinetics.25 Thirdly, the active materials achieve enhanced distribution across the substrate surfaces, which can expose more active sites, significantly improving the water-splitting performance.26 Finally, circumventing the use of binders mitigates potential binder degradation, thereby enhancing electrode stability.27 Over the years, there has been a noticeable upward trend in the number of in situ grown TMSs used as electrocatalysts. The increasing focus on electrocatalysts, as demonstrated in Fig. 1c, indicates a recognition within the scientific community of the potential of in situ grown TMSs to overcome the challenges inherent in energy conversion and storage. The schematic comparison between in situ grown and ex situ grown electrodes is illustrated in Fig. 1b.

In this review, we comprehensively examine recent advancements in the design, synthesis, and application of in situ grown TMSs for electrocatalytic water splitting. An overall introduction to electrodes is provided, including the active TMS materials, substrates and the preparation strategies. Furthermore, we also discuss the underlying mechanisms that contribute to their superior catalytic activities and highlight the role of structural and compositional modifications in optimizing their performance. By delving into the current challenges and potential solutions, this review aims to provide a holistic understanding of the advancements in TMSs as electrocatalysts. Our goal is to shed light on the future directions and opportunities in this rapidly evolving field, ultimately contributing to the realization of efficient and sustainable hydrogen production through EWS.

2. Overview of in situ grown TMS-based working electrodes for EWS

The selection of appropriate active materials is critical for optimizing the electrochemical performance of the electrodes, as these materials directly influence the efficiency and stability of the electrochemical reactions. Equally important is the choice of electrode substrates, which serve as the foundational support for the active materials, ensuring effective electron transfer and mechanical integrity during operation. Various fabrication techniques are employed to in situ grow these materials onto the substrates, with each method tailored to enhance the overall electrode architecture and improve its functional properties. This strategic combination of carefully selected materials and precise fabrication methods is essential for advancing the development of high-performance electrodes across various applications, including energy storage and catalytic processes. Therefore, in this section, a comprehensive overview of the key components is involved, including the active materials, electrode substrates, and fabrication methods.

2.1 TMS synopsis

TMSs are a class of compounds composed of transition metals and sulfur. These materials are characterized by their diverse crystal structures and chemical compositions, which result from the ability of transition metals to exhibit multiple oxidation states and coordinate with sulfur atoms in various ways. The history of TMS materials spans over a century. As early as 1923, Linus Pauling and colleagues reported the crystal structure of MoS2.28 In 1969, Wilson et al. provided a systematic introduction to transition metal dichalcogenides, detailing their optical, electrical, and structural properties.29 The structures of TMS materials can be broadly classified into two categories: (1) layered structures and (2) non-layered structures. Transition metal dichalcogenides (TMDs), such as MoS2 and tungsten disulfide (WS2), belong to the layered structure category and are typically represented as MS2 (where M stands for Mo, W, etc.). The remaining TMS materials, which include various transition metals like cobalt (Co), iron (Fe), and nickel (Ni), fall into the non-layered structure category and are generally denoted as MxSy.30

In single layers of TMS materials, the transition metal (M) is sandwiched between two layers of chalcogens (S) to form an S–M–S trilayer structure. These single sheets are held together by van der Waals forces, forming bulk materials with layered structures.31 Within these layered structures, variations in surface orientation lead to basal and edge planes with anisotropic properties.32 Depending on the coordination geometry, layered TMS materials exhibit two primary phases, corresponding to octahedral and trigonal prismatic coordination. Each single layer in multilayered TMS materials can adopt either of these phases, resulting in structures with different coordination styles.33 However, three of the most commonly observed structures are 1T, 2H, and 3R, as illustrated in Fig. 2a.34 For instance, WS2 and MoS2 predominantly exhibit the 2H phase, which is the most thermodynamically stable.38 Nonetheless, phase transitions, such as from 2H to 1T, can occur through ion intercalation, a process often used to modify electrochemical performance.


image file: d4ta06197g-f2.tif
Fig. 2 (a) Unit cell structure of 1T MoS2, 2H MoS2 and 3R MoS2; reproduced with permission from ref. 34. Copyright 2015 ACS. (b) Crystal structures of FeS2; reproduced with permission from ref. 35. Copyright 2020 ACS. (c) Crystal structures of nickel sulfides; reproduced with permission from ref. 36. Copyright 2020 RSC. (d) Schematic diagram of three different types of substrates used for in situ grown TMS electrodes; reproduced with permission from ref. 37. Copyright 2023 MDPI. (e) Schematic diagram of the hydrothermal strategy for the preparation of in situ grown TMS electrodes. (f) Diagram of the electrodeposition strategy for the preparation of in situ grown TMS electrodes.

In contrast, non-layered TMS materials typically adopt either the pyrite or marcasite structure. In these structures, transition metals like Fe, Co, Ni, copper (Cu), and zinc (Zn) are coordinated with sulfur in an octahedral arrangement.39 For instance, Fig. 2b illustrates the pyrite and marcasite structures of iron sulfides (FeS2). In the pyrite structure, iron atoms, depicted as yellow spheres, occupy the corners and face centers of the cubic unit cell, while sulfur atoms, shown as orange spheres, form S2 (disulfur) dimers aligned along the crystallographic axes. The pyrite structure is characterized by its high symmetry, with each iron atom octahedrally coordinated by six sulfur atoms. In contrast, the marcasite structure, while maintaining the octahedral coordination between iron and sulfur atoms, exhibits a more complex lattice arrangement. The FeS2 units are connected differently compared to the pyrite structure, leading to a lower symmetry in the marcasite structure. The S2 dimers in marcasite are more closely spaced and arranged in a manner that gives the structure its distinct, less symmetrical configuration.35 Beyond pyrite and marcasite structures, TMSs can also exhibit other structural forms. For example, Fig. 2c depicts the crystal structures of various polymorphs of nickel sulfides, including NiS, NiS2, Ni3S2, and Ni9S8. Each polymorph demonstrates distinct geometric arrangements and coordination environments of Ni and sulfur (S) atoms, as visualized through the depicted polyhedron and atomic networks. Each polymorph demonstrates distinct geometric arrangements and coordination environments of Ni and sulfur (S) atoms, as visualized through the depicted polyhedron and atomic networks.36 The properties of non-layered TMSs vary depending on the d-orbitals of the transition metals. For instance, NiS2 exhibits poor conductivity, whereas CoS2 demonstrates metal-like conductivity.

2.2 The substrate materials for the preparation of in situ grown TMS electrodes

As illustrated in Fig. 2d, a wide range of substrates have been explored for the development of in situ grown catalysts. These substrates include carbon-based materials such as carbon fiber (CF), carbon cloth (CC), graphite, graphene, reduced graphene oxide (rGO), and carbon nanotubes (CNT);40,41 metal-based materials like nickel foam (NF) and copper wire (CW);42,43 and other conductive materials such as fluorine-doped tin oxide (FTO) and indium tin oxide (ITO). These materials not only address the conductivity challenges associated with traditional binders but also pave the way for new approaches in the design and application of TMSs to enhance the efficiency and durability of EWS technologies.

Carbon-based materials are particularly well-suited as substrates for TMS catalysts due to their high conductivity, outstanding stability, and cost-effectiveness. The in situ growth of TMSs on these carbon substrates substantially enhances catalytic activity by providing an efficient pathway for charge transfer.44,45 CF and CC, in particular, are distinguished by their excellent mechanical and electrical properties, attributable to their unique preparation methods and high carbon content, making them highly effective in catalytic applications. Graphite, a naturally occurring crystalline form of carbon, is commonly used in lithium-ion batteries (LIBs) and is known for its unique structure and properties. It is one of the most stable forms of carbon under standard conditions, making it valuable in environments requiring stability under corrosive conditions. With its 2D-layered structure, graphite is an excellent conductor of electricity due to the mobility of free electrons within its layers, making it ideal for use in electrodes for batteries and catalysts.46 Graphene, characterized by its single layer of carbon atoms arranged in a hexagonal lattice, is the thinnest material known and exhibits exceptional electrical properties.47 Similarly, carbon nanotubes (CNTs), which are composed of carbon atoms arranged in a hexagonal pattern and rolled into tubular structures, offer a very high surface area due to their nanoscale dimensions, making them particularly beneficial for catalytic applications. rGO, produced by reducing graphene oxide, retains a combination of conductive graphene-like regions and residual oxygen functionalities. This hybrid structure, resulting from the partial restoration of the sp2 carbon network, enhances its effectiveness as a substrate for various applications.48

Metal-based substrates also play a crucial role in electrocatalysis. These substrates are chosen for their excellent conductivity, chemical stability, and ability to support active catalytic materials. NF, with its unique 3D porous structure, outstanding conductivity, and high surface area, not only serves as an excellent support for TMSs but also possesses inherent catalytic activity. Its interconnected network of nickel strands provides ample sites for catalyst loading.49,50 Cu substrates are commonly used in the form of copper foil or CW. Cu's high conductivity and ease of fabrication make it a popular choice for electrocatalytic applications.51 Other metal-based substrates, such as stainless steel and titanium, are also utilized; however, their relatively poor surface activity and higher cost often limit their use compared to Ni and Cu-based substrates.52 The combination of conductivity, chemical stability, and structural support makes metal-based substrates essential for enhancing the performance and longevity of catalysts in water splitting and other electrochemical processes. These substrates not only facilitate the necessary electrochemical reactions but also provide a foundation for the development of advanced, cost-effective energy conversion technologies.

FTO and ITO are two widely used transparent conducting oxides that serve as effective substrates for catalysts in various electrochemical and photoelectrochemical applications. FTO is composed of tin oxide (SnO2) doped with fluorine atoms, which enhances its electrical conductivity while maintaining high transparency in the visible spectrum. The doping process introduces free electrons into the tin oxide matrix, improving its conductive properties without significantly compromising its transparency.53,54 ITO is a ceramic material composed of indium oxide (In2O3) and SnO2, typically containing 90% In2O3 and 10% SnO2. This combination results in a material with high electrical conductivity and excellent transparency across the visible light spectrum. ITO is also chemically stable and can withstand harsh conditions, making it a robust substrate in many catalytic applications.55,56

2.3 Preparation strategies for in situ grown TMS materials

In the preparation of in situ grown electrodes, the active materials are directly grown on the surface of substrates without the use of binders and other conductive additives. Various methods, including hydrothermal, solvothermal,57–59 electrodeposition,60,61 and chemical deposition,62 have been widely employed to synthesize in situ grown catalysts. Among these, hydrothermal and electrodeposition strategies are particularly noteworthy for their ease of modification, straightforward operational procedures, and robust stability.

Hydrothermal is the predominant method for preparing in situ grown samples. This process utilizes water as the medium, where nucleation and conformal growth of TMS materials occur on the defects of substrates. The diagram of in situ grown TMS electrodes prepared via the hydrothermal method is shown in Fig. 2e. By adjusting preparation parameters such as temperature and pH value, various phases and morphologies of samples can be synthesized.63 For instance, in 2019, Qu et al. synthesized a series of phosphorus-doped cobalt sulfides on carbon cloth by merely adjusting the hydrothermal temperature (160 °C, 180 °C, and 200 °C). The XRD results indicated that three different phases of cobalt sulfides were synthesized (P-CoS2/Co1−xS, P-Co1−xS, and P-Co9S8).64 Additionally, the high temperature and pressure during the preparation process provide substantial energy, facilitating the doping or insertion of other elements. Despite these advantages, the hydrothermal method sometimes requires combination with other techniques, such as thermal treatment, sulfurization, or an additional hydrothermal step, to achieve effective sample preparation.65,66

Compared to the hydrothermal method, electrodeposition involves simpler processes and requires less preparation time and energy. This method is typically conducted in a three-electrode system, where the selected substrate serves as the working electrode, and the target materials are directly deposited onto its surface, accompanied by electron transfer (Fig. 2f). By adjusting the ratio of raw materials and deposition durations, electrodes with varying loading masses, morphologies and structures can be synthesized.67 For example, Zeng et al. utilized a simple potentiostatic electrodeposition method to prepare a NiSx catalyst on copper wire (NiSx/CW), with the morphology varying according to the ratio of Ni to S.68 Although doping is generally challenging to achieve via electrodeposition due to the lower energy input, ion intercalation can be readily facilitated by modifying the potential or current during the process. This technique is commonly employed to design novel electrodes with unique structures and outstanding electrochemical performance.

Hydrothermal and electrodeposition methods are the primary strategies for the preparation of in situ grown TMS electrodes and have been extensively discussed. However, other methods, such as thermal sulfidation, solvothermal, and template-assisted methods also play crucial roles in the preparation of in situ grown TMS catalysts. These alternative strategies offer unique advantages, including precise control over nanostructures and compositions, as well as the ability to produce catalysts with high surface areas, uniform pore distributions, and complex functional geometries. The advancement of these methodologies represents a significant evolution from traditional synthesis approaches, enabling the development of catalysts that are more active, selective, and durable.

3. Boosting the performance of electrocatalysis via in situ grown TMSs

In the preparation of in situ grown TMS catalysts, the active samples are directly grown on substrates, primarily serving as the key catalysts for overall water splitting.69 As a result, the properties of these active materials have a significant impact on catalytic performance. Compared to TMOs, TMSs exhibit better electrical conductivity due to their narrow band gap. Moreover, the abundant active sites in TMSs can efficiently adsorb hydrogen and hydroxide ions from electrolytes, making them ideal materials for EWS.70–72 For instance, in Zhang et al.’s research, a TMO-based material, MnCo2O4 was initially selected as a catalyst for the OER process. In LSV tests, the MnCo2O4/TM electrode required a high overpotential of 693 mV to reach a current density of 50 mA cm−2, indicating substantial energy consumption. Subsequently, they used a sulfurization method to prepare a MnCo2S4/TM electrode, which showed a significant improvement, reducing the overpotential from 693 mV to 325 mV. Additionally, the Tafel slope improved from 228 mV dec−1 to 115 mV dec−1.73 Similarly, in 2020, Joyner compared the HER performance of MoO2/MoO3 and MoS2, both of which were prepared via ex situ growth strategies. The onset potential of MoS2 for the HER was around −200 mV, representing a significant improvement compared to MoO2/MoO3 (−598 mV).74 However, the −200 mV onset potential of MoS2 still cannot meet the requirement for practical applications. In 2021, Dong and his team employed an in situ growth strategy to prepare a 1T MoS2/rGO composite. This in situ grown composite demonstrated an exceptionally low Tafel slope of 56.7 mV dec−1 and around 100 mV onset potential.75

These studies highlight that TMS-based materials possess superior catalytic activities compared to TMO-based materials. Moreover, the adoption of in situ growth techniques can significantly enhance EWS performance. Therefore, the development of in situ grown TMS catalysts holds great promise for advancing electrochemical and catalytic technologies.

3.1 In situ grown molybdenum sulfides for enhancing the electrocatalytic HER

MoS2 is generally considered to be more effective for the HER rather than for the OER. The electronic structure of MoS2 is well-suited for facilitating the adsorption and subsequent reduction of protons, which is crucial for the HER. During the OER, MoS2 can undergo oxidation, which compromises its structural integrity and reduces its catalytic performance over time. This makes it less stable under the harsh oxidative conditions of the OER compared to the HER, where such degradation is less of a concern. Despite this, researchers continue to investigate ways to improve its OER performance through doping, phase engineering, and other modifications. For instance, Zhao et al. composited MoS2 with amorphous CoFeOx(OH)y and grew it in situ on carbon paper by the hydrothermal method for the preparation of high efficiency OER catalysts.76 While these modifications can improve MoS2 for OER applications, its inherent properties suggest that selecting alternative materials may currently be the most effective approach for OER catalysis.
3.1.1 In situ grown 2H-MoS2 electrodes. MoS2 can exist in three distinct phases, 2H, 1T and 3R, depending on its coordination structure. The 2H phase is the most stable structure for MoS2; however, its poor conductivity and inert catalytic sites significantly hinder its application in EWS. To overcome these limitations, in situ growth of 2H-MoS2 on conductive substrates has been explored as an effective strategy. In 2016, Tang et al. successfully synthesized a 2H-MoS2/N-rGO nanocomposite through a two-step process. Initially, 2H-phase MoS2 was prepared and then combined with N-rGO via a chemical deposition method. This approach ensured a uniform distribution of MoS2 on the N-rGO substrate, effectively creating a robust, interconnected network that promotes efficient electron transport and catalytic activity. The resulting MoS2/N-rGO-180 nanocomposite demonstrated excellent HER activity with an onset potential of just −5 mV and a small Tafel slope of 41.3 mV dec−1. Additionally, cycle stability tests confirmed that the nanocomposite maintained its structural integrity and catalytic performance over 5000 CV cycles between −50 and −150 mV, indicating its suitability for long-term electrocatalytic applications. This research also intentionally increased the interlayer spacing within the MoS2 structure (Fig. 3a), facilitating greater accessibility of catalytic sites and improving HER kinetics.77 A similar enhancement was also observed by Huang et al. in 2020, who synthesized 2H-MoS2 on graphite felt (GF) using a unique in situ growth approach. GF served a dual purpose as both the growth substrate for MoS2 nanoflowers and a component of the composite catalyst. The as-prepared MoS2/GF composite exhibited remarkable catalytic activity, maintaining stable HER performance even after 3000 cyclic voltammetry scans, indicating both high activity and durability (Fig. 3b). The promoted electron transfer at the MoS2/GF electrodes, clearly displayed in Fig. 3d was a key factor contributing to this excellent catalytic performance.45
image file: d4ta06197g-f3.tif
Fig. 3 (a) The schematic of enlarged interlayer space of MoS2; reproduced with permission from ref. 77. Copyright 2016 Wiley. (b) Polarization curves of MoS2/GF activated at different voltages (−0.8 V, −1.2 V and −1.0 V), and polarization curves of MoS2/GF before and after 3000th cycles; reproduced with permission from ref. 45. Copyright 2020 Elsevier. (c) Schematic diagram of the synthesis process of both the MoS2/graphene superlattice and the control superlattice; reproduced with permission from ref. 78. Copyright 2018 ACS. (d) Schematic of the electron transfer route at the catalysts during the HER process; reproduced with permission from ref. 45. Copyright 2020 Elsevier. (e) Schematic illustration of the preparation of MoSx/carbon cloth and the contact measurement result before and after H2 plasma treatment; reproduced with permission from ref. 79. Copyright 2016 Wiley. (f) The schematic illustration of the mechanism of the HER process for the as-prepared 5.2% Rh–MoS2 catalyst; reproduced with permission from ref. 80. Copyright 2017 Wiley. (g) The Gibbs free energy of hydrogen evolution for 2H-MoS2, 1T-MoS2 and V/1T-MoS2; reproduced with permission from ref. 81. Copyright 2021 Elsevier.

As previously mentioned, MoS2 has two types of catalytic sites, located on the basal plane and at the edge. The basal plane sites exhibit inert catalytic performance due to their low surface energy, while the edge sites are highly active. Therefore, strategies such as surface engineering or creating defects to expose more edge sites can further enhance the catalytic activity of in situ grown 2H-MoS2. Lu et al. presented an innovative approach to enhance the catalytic efficiency of amorphous MoSx on carbon cloth for the HER by creating high-density sulfur vacancies using remote H2 plasma. This modification led to an increased density of active sites and improved surface hydrophilicity, significantly boosting catalytic activity and durability under high current densities. The preparation schematic and hydrophilicity test results are shown in Fig. 3e. The overpotential of a-MoSx decreased by about 63 mV after plasma treatment, attributed to the increased active site density.79

Other strategies have also been implemented to modify the electrochemical performance of in situ grown 2H-MoS2 electrodes. For example, doping is widely used to improve the catalytic, electrical, and structural characteristics of materials, particularly in catalysts for energy conversion and storage. In 2017, Cheng et al. employed a solvothermal process to synthesize Rh-doped 2H MoS2 catalysts to enhance HER performance. 2H-MoS2 nanosheets, prepared via a lithium intercalation method, serve as the base onto which Rh nanoparticles are deposited using ethylene glycol at 90 °C. This synthesis approach resulted in a Rh–MoS2 composite with exceptional HER performance, achieving a Tafel slope of only 24 mV dec−1, attributed to the hydrogen spillover effect and the synergistic interaction between Rh and MoS2; the mechanism underlying these effects is illustrated in Fig. 3f.80 Another study by Qian et al. in 2021 demonstrated the enhancement of MoS2's electrochemical performance through zinc doping and the in situ growth of MoS2 on reduced graphene oxides. The synthesized Zn-doped 2H-MoS2 nanosheets on rGO (Zn–MoS2-rGO) showcased superior HER performance, highlighted by a low Tafel slope of 69 mV dec−1, with the improvement primarily attributed to enhanced conductivity.44 In addition to surface engineering and elemental doping, intercalation can also efficiently improve the electrochemical performance by modifying the interlayer space. In 2022, our group introduced a facile electrodeposition method to produce in situ grown Ni-modified MoS2. This innovative approach resulted in 2H-MoS2 with increase interlayer spacing due to Ni intercalation, showcasing excellent electrocatalytic activity with an onset potential of 139 mV and a Tafel slope of 62 mV dec−1, attributed to the synergistic effect of the Ni–Mo–S composition.82

3.1.2 In situ grown 1T-MoS2 electrodes. The 1T phase of MoS2 offers several advantages over the 2H phase, making it particularly attractive for catalytic applications. Firstly, 1T-phase MoS2 is metallic, providing significantly higher electrical conductivity than the semiconducting 2H phase. Secondly, the 1T phase features a distorted octahedral structure, which disrupts symmetry and creates more edge sites, thereby increasing the number of active sites. Thirdly, the 2D structure of 1T-MoS2 typically results in a higher surface area compared to 2H-MoS2, further enhancing the availability of active sites for chemical reactions and improving catalytic efficiency. Consequently, the selection of 1T-MoS2 as the active material for the preparation of in situ grown electrodes has garnered significant attention in recent years.

In 2018, Xiong et al. introduced a single-layered MoS2/graphene superlattice fabricated through a facile solution-phase restacking method. This novel preparation method, detailed in Fig. 3c, employed chemical exfoliation via Li intercalation to induce metallic properties in MoS2 nanosheets, thereby improving charge transfer kinetics and catalytic activity. The resulting structure exhibited outstanding electrochemical properties as an HER catalyst, with an 88 mV onset potential and a Tafel slope of 48.7 mV dec−1, highlighting the potential of two-dimensional nanosheets in advanced energy applications.78 Further modifications have also been explored to improve the catalytic activity of 1T-MoS2. In 2021, our group utilized vanadium (V) to fabricate in situ grown V-doped 1T MoS2 catalysts on carbon paper via a straightforward one-step hydrothermal technique. The resulting V-doped 1T-MoS2 exhibited exceptional HER performance, with a Tafel slope of just 54 mV dec−1 and an onset potential of 102 mV. Additionally, during a 12-hour constant current test, the V-doped 1T-MoS2 demonstrated remarkable stability with only a minimal increase in overpotential. The Gibbs free energy of hydrogen adsorption (ΔGH), calculated using density functional theory (DFT) and shown in Fig. 3g, indicated that V-doped 1T-MoS2 had the lowest ΔGH (0.03 eV), underscoring its superior HER performance.81

However, the metallic 1T phase, despite its advantages of high conductivity and increased active sites, is thermodynamically less stable than the semiconducting 2H phase. This instability can result in a spontaneous phase transition, particularly under operational conditions, which diminishes the material's catalytic efficiency over time. Furthermore, the synthesis of 1T-MoS2 typically requires complex procedures, such as chemical exfoliation or intercalation methods, making it challenging to scale up for industrial applications. Maintaining the 1T phase during the synthesis and operational stages is thus a significant obstacle. Researchers are actively exploring methods to stabilize the 1T phase, such as doping or using specific substrates, but achieving long-term stability remains a critical challenge that must be addressed before 1T-MoS2 can be widely adopted in practical catalytic applications. The 3R phase of MoS2 possesses a rhombohedral structure, similar to the 2H phase, but with a distinct stacking sequence. While it also exhibits semiconducting properties, the 3R phase is less commonly studied compared to the 2H phase. Its rhombohedral stacking and semiconducting nature result in lower catalytic activity compared to the metallic 1T phase, which provides higher conductivity and more abundant active sites. Consequently, the 3R phase has received less attention in research for applications that demand high conductivity and efficient catalysis.83 The comprehensive comparison of the catalytic activities of the in situ grown molybdenum sulfide catalysts discussed in this study is shown in Table 2.

Table 2 Comparison of EWS performance of in situ grown molybdenum sulfide catalysts reported in the literature
Catalyst Substrate Preparation Overpotential η (mV) at 10 mA cm−2 Tafel slope (mV dec−1) Electrolyte Ref.
HER HER
Defective-MoS2/rGO rGO Hydrothermal 154.77 56.17 0.5 M H2SO4 75
Annealed-MoS2/rGO rGO Hydrothermal 278.04 128.9 0.5 M H2SO4 75
Zn–MoS2-rGO rGO Hydrothermal 300 (169 mA cm−2) 69 0.5 M H2SO4 44
MoS2/N-RGO-180 rGO Hydrothermal 56 41.3 0.5 M H2SO4 77
a-MoSx/CC Carbon cloth Hydrogen plasma 143 39.5 0.5 M H2SO4 79
V-doped 1T MoS2/CP Carbon paper Hydrothermal 102 (onset potential) 54 0.5 M H2SO4 81
MoS2/graphene Graphite Chemical exfoliation 137 48.7 0.5 M H2SO4 78
MoS2/GF(−1.0 V) Graphite felt Hydrothermal 82 48 0.5 M H2SO4 45
Ni0.05Mo0.95S2/NF Nickel foam Electrodeposition 215 62 0.5 M H2SO4 82


3.2 In situ grown cobalt sulfides for enhancing EWS

Cobalt sulfides are a class of inorganic compounds composed of cobalt and sulfur, with the general formula CoxSy, where x and y vary depending on the specific phase or stoichiometry of the compound. They exist in several phases, including cobalt monosulfide (CoS), cobalt disulfide (CoS2), cobalt pentlandite (Co9S8), and cobalt trisulfide (Co3S4). Among them, CoS2 exhibits the best electrical conductivity and abundant active sites, making it particularly valuable in electrocatalysis. Co9S8, with its cubic crystal structure, is more complex, featuring cobalt atoms occupying multiple distinct sites within the lattice. While it is a conductor, its conductivity is generally lower than that of CoS2; however, its unique structure supports various electronic applications where moderate conductivity is required. CoS and Co3S4 are semiconductors, but their conductivity can be tailored through doping or the creation of nanostructures to enhance their performance in electronic and catalytic applications.

The in situ growth of cobalt sulfides on substrates results in a well-integrated interface that facilitates efficient electron transfer. This close interaction between the cobalt sulfide catalyst and the conductive substrate minimizes resistance and improves the overall catalytic performance, particularly in reactions such as the HER and OER. Therefore, the following section will first focus on in situ grown CoS2, highlighting its exceptional conductivity and abundant active sites. Subsequently, other in situ grown cobalt sulfides, including CoS, Co9S8, and Co3S4, will be discussed in relation to their various valence states and catalytic properties.

3.2.1 In situ grown CoS2 electrodes. The metallic nature of CoS2 provides excellent conductivity, while its unique crystal structure offers numerous active sites for hydrogen adsorption and reduction. These features make in situ grown CoS2 a promising catalyst for the HER, with the potential to replace more expensive noble metal catalysts in practical applications.

In 2013, Kong et al. discussed first-row transition metal dichalcogenides as novel, non-precious catalysts for the HER in acidic electrolytes. Their preparation method involved conformally coating carbon black nanoparticles with a cobalt layer as the precursor, followed by sulfurization in a tube furnace. By utilizing this facile process for growing polycrystalline dichalcogenide films, the study opened new avenues for developing efficient catalysts for energy technologies, particularly in applications requiring acid stability, such as proton exchange membrane electrolysis units. The electrochemical measurements revealed that Fe0.43Co0.57S2 core–shell nanoparticles excelled in HER efficiency, ranking among the top non-precious material-based catalysts. These nanoparticles notably outperform their Fe0.43Co0.57S2 film counterparts, achieving higher current densities at substantially lower overpotentials. Furthermore, the similarity in Tafel slopes between the nanoparticles and the films suggested similar surface chemical behaviours during the HER process, as detailed in Fig. 4a.84 In 2014, Faber et al. synthesized CoS2 on graphite disk substrates using three different methods, resulting in various morphologies: CoS2 films, CoS2 microwires (CoS2 MWs) and CoS2 nanowires (CoS2 NWs), as shown in Fig. 4b. SEM images indicate that these three cobalt sulfides were intimately grown on the graphite substrate, enhancing their mechanical stability. Among them, CoS2 NWs exhibited the most impressive HER performance, attributed to the facilitated release of gas bubbles generated at the electrode surface, which underscores the advantages of micro- and nano-structural optimization in catalyst design. CoS2 NWs required only 145 mV overpotential to achieve a current density of 10 mA cm−2, which is lower than that of CoS2 films and CoS2 MWs (190 mV and 158 mV), and the Tafel slope of CoS2 NW was just 51.6 mV dec−1 (Fig. 4c).85


image file: d4ta06197g-f4.tif
Fig. 4 (a) Polarization curves of the as-prepared catalysts grown on glassy carbon electrodes for the HER; reproduced with permission from ref. 84. Copyright 2013 RSC. (b) Preparation schematic of CoxSy/ACF by a simple hydrothermal method; reproduced with permission from ref. 85. Copyright 2014 ACS. (c) Polarization curves of the as-prepared catalysts for the HER process; reproduced with permission from ref. 85. Copyright 2014 ACS. (d) Preparation diagram of in situ grown CoS2@Co nanoparticles on the CNT substrate; reproduced with permission from ref. 86. Copyright 2018 Elsevier. (e) Detailed preparation schematic for FeS2-MoS2@CoS2-MOF; reproduced with permission from ref. 87. Copyright 2022 Elsevier.

In situ grown CoS2 has also demonstrated significant potential as an OER catalyst, largely due to its ability to provide a high density of active sites. The in situ growth process ensures a robust connection between CoS2 and the substrate, which not only facilitates efficient electron transfer but also significantly enhances the overall durability of the catalyst. As a result, in situ grown CoS2 emerges as a promising candidate for overall water splitting, effectively catalysing both the HER and OER. Its bifunctional nature, coupled with cost advantages and the potential for further optimization, positions it as an attractive material for sustainable hydrogen production. In 2018, Wang et al. employed a novel preparation strategy that involved carbonization followed by vulcanization to create S and N co-doped carbon nanotube-encapsulated core–shelled CoS2@Co nanoparticles. This structure, coupled with the Mott–Schottky effect between metallic cobalt and semiconductive cobalt disulfide, significantly boosted catalytic efficiency for both the HER and OER, offering a promising low-cost, high-activity, and stable alternative for water splitting applications (Fig. 4d).86 In 2022, Chhetri et al. developed a novel catalyst for water splitting, featuring bimetallic sulfide-coupled mesoporous CoS2 nanoarray hybrids derived from a 2D MOF. As shown in Fig. 4e, the electrocatalyst was synthesized through a method involving the growth of a 2D MOF on NF, followed by annealing with sulfur to form CoS2 nanoarrays. These were then coupled with FeS2@MoS2 layers via a hydrothermal process, resulting in a structure with enhanced electrocatalytic activity due to its abundant heterointerfaces, mesoporosity, and multimetallic active centres. This innovative synthesis method yielded a material that demonstrates superior activity in both the HER (92 mV overpotential at 10 mA cm−2 current density) and OER (211 mV overpotential at 20 mA cm−2 current density).87

3.2.2 Other in situ grown cobalt sulfide-based electrodes. In addition to the well-studied CoS2, other in situ grown cobalt sulfides have also attracted considerable attention in the field of electrocatalysis. By leveraging the versatility of cobalt sulfides, researchers have developed a range of catalysts that offer promising alternatives to more expensive or less abundant materials.

In situ grown CoS catalysts are particularly valued for their structural integrity and resistance to degradation under operating conditions, making them highly effective for sustainable water-splitting applications. In 2018, Kale et al. prepared a binder-free cobalt sulfide thin film material on stainless steel via a chemical bath deposition (CBD) approach. The as-prepared CoS/SS displayed satisfactory conductivity due to its binder-free construction and thin-film structure. Compared to other cobalt sulfides, CoS/SS required lower kinetic energy in the OER process, with a Tafel slope of only 55 mV dec−1 in an alkaline medium.88 Subsequently, in 2020, Wang et al. employed a two-step electrodeposition method (Fig. 5a) to synthesize a novel CoS@NiFe/Ni foam heterostructure catalyst, creating an amorphous NiFe layer coated with a CoS film on commercial NF. This structure exhibited exceptional OER performance, requiring a low overpotential of 175 mV to drive a current density of 10 mA cm−2 and demonstrating high stability in 1 M KOH alkaline solutions. The outstanding performance was attributed to the strong electronic interactions between the CoS film and the NiFe layer, which regulated redox states and promoted electrocatalytic activity (Fig. 5b).89


image file: d4ta06197g-f5.tif
Fig. 5 (a) Schematic illustration of the preparation of CoS@NiFe/NF by the electrodeposition method; reproduced with permission from ref. 89. Copyright 2020 ACES. (b) Polarization curves of the as-prepared catalysts for the OER; reproduced with permission from ref. 89. Copyright 2020 ACES. (c) Preparation diagram of edge-MoS2/Co3S4@NFs; reproduced with permission from ref. 90. Copyright 2020 Elsevier. (d) Polarization curves of the as-prepared catalysts on carbon fibre for the OER; reproduced with permission from ref. 91. Copyright 2020 Wiley. (e) Intermediate adsorption configurations of *H and *OH on the NiFe LDH and NiCo2S4@NiFe LDH heterostructure surfaces, respectively; and the electron transfers at the interface between NiCo2S4 and NiFe LDH; regions of charge accumulation and depletion are depicted in yellow and light blue, respectively; reproduced with permission from ref. 92. Copyright 2017 ACS. (f) LSV curves of the as-prepared catalysts; reproduced with permission from ref. 93. Copyright 2019 Elsevier. (g) Preparation schematic of FeCo2S4/NF and their corresponding morphologies; reproduced with permission from ref. 94. Copyright 2018 ACS. (h) Preparation schematic of Zn doped and un-doped Co9S8 catalysts; reproduced with permission from ref. 95. Copyright 2020 Elsevier.

Co3S4, a cobalt sulfide compound with a spinel-type crystal structure, features a cubic lattice where cobalt atoms occupy both tetrahedral and octahedral sites, while sulfur atoms form a close-packed framework. This unique arrangement contributes to the material's distinct electrochemical properties, making it an attractive catalyst for the HER. In 2020, Peng et al. combined Co3S4 nanorods with MoS2 nanoflakes via a multi-step preparation process involving the hydrothermal growth of Co3S4 nanorods on nickel foam, followed by decoration with MoS2 nanoflakes (Fig. 5c). This arrangement maximized the exposure of catalytically active edges, significantly enhancing HER performance (η10 = 90.3 mV and η1000 = 502.0 mV). The enhanced performance was attributed to the effective distribution of active sites and the synergistic effect of the Co3S4 and MoS2 components. The hierarchical heterostructure, combined with NF, ensured optimal electron transfer throughout the catalyst, contributing to its enhanced electrocatalytic efficiency. The as-prepared edge-MoS2/Co3S4@NFs maintained its performance over 20 hours at 100 mV current density without significant loss in activity, highlighting its potential for long-term applications.90

Co3S4 features cobalt in mixed valence states (Co2+ and Co3+), which enhances redox activity and facilitates efficient electron transfer during catalytic reactions, particularly beneficial for the OER, where multiple electron transfers are required. Thangasamy et al. doped the Fe element into cobalt sulfides and grew the active samples on carbon fibre using a simple hydrothermal method. The standard reference pattern of Co3S4 was observed in PXRD results. The resulting Fe-CoxSy/ACF demonstrated a remarkably low overpotential (304 mV) at a constant current density (10 mA cm−2) and a small Tafel slope of 54.2 mV dec−1 in an alkaline medium for the OER (Fig. 5d). The enhanced catalytic performance was primarily attributed to the 3D ferrocene-cobalt sulfide nanoarchitectures, which featured a high concentration of vertical nanosheets uniformly distributed with nanoparticles.91 Consequently, Co3S4 has gained attention as a highly effective bifunctional catalyst for both the HER and OER. In 2015, Danni Liu et al. successfully prepared NiCo2S4 nanowires on carbon cloth (NiCo2S4 NA/CC) via a straightforward hydrothermal process, followed by sulfidation treatment. In the as-prepared NiCo2S4 NA/CC, Co exhibited the coexistence of Co2+ and Co3+. The homogeneous nanostructure of NiCo2S4 NA/CC displayed excellent electrochemical activity for both the HER and OER, characterized by low overpotential (305 and 340 mV) at a current density of 100 mA cm−2.96 Similarly, in 2017 Jia Liu et al. employed a three-step hydrothermal method to synthesize a novel hierarchical NiCo2S4@NiFe LDH/NF catalyst, which exhibited outstanding HER and OER performance. The Co elements in NiCo2S4@NiFe LDH/NF also showed the coexistence of +2 and +3 values. The as-prepared catalysts displayed a low overpotential of 200 mV to drive a current density of 10 mA cm−2 and exceptional stability for the HER, while the OER performance was also excellent, with only 201 mV overpotential at a current density of 60 mA cm−2. DFT calculations demonstrated a strong synergistic effect between the NiCo2S4 and NiFe LDH components, which is beneficial for adjusting the interfacial electronic structure, as shown in Fig. 5e. Moreover, the overall water-splitting voltage of NiCo2S4@NiFe LDH/NF was only 1.6 V at 10 mA cm−2, making it a promising material for efficient and cost-effective electrocatalysis in water-splitting applications.92 In 2018, Hu et al. adopted a two step strategy to synthesize FeCo2S4 on a nickel foam substrate (FeCo2S4/NF). According to XRD results, the main phase of the as-prepared catalysts was Co3S4. Initially, FeCo2O4/NF, as the precursor material, was prepared by a hydrothermal method. This precursor was then annealed at 160 °C for 8 hours to convert it into FeCo2S4/NF. The detailed preparation process is depicted in Fig. 5g. FeCo2S4/NF demonstrated outstanding HER and OER properties in a strong alkaline electrolyte.94

Co9S8, another cobalt sulfide compound, features a unique crystal structure where cobalt atoms are arranged in both octahedral and tetrahedral sites within a sulfur-rich lattice. This configuration provides a diverse range of active sites, enhancing the material's catalytic performance. Karikalan and his group first prepared CoMoO4/SDC precursors through a standard chemical reaction, and then CoMoS4/SDC was synthesized via a hydrothermal strategy. According to the characterization results, Co9S8 was the main phase in CoMoS4/SDC, and the formation of Co9S8 on the sulfur-doped carbon (SDC) surface, known for its advantageous role in the HER from prior studies, was observed. Compared with individual SDC or Co9S8, CoMoS4/SDC exhibited better HER performance (Fig. 5f), which is attributed to the enhancement of ion exchange due to the interaction between Co2+ and sulfur ions within SDC. According to the characterization results, Co9S8 was the main phase in CoMoS4/SDC, and the formation of Co9S8 on the sulfur-doped carbon (SDC) surface, known for its advantageous role in the HER from prior studies, was observed. Compared with individual SDC or Co9S8, CoMoS4/SDC exhibited better HER performance (Fig. 5f), which is attributed to the enhancement of ion exchange due to the interaction between Co2+ and sulfur ions within SDC. During long-term cycling in 0.5 M H2SO4 at a high current density (220 mV cm−2), CoMoS4/SDC exhibited notable stability compared to unsupported CoMoS4, showing only a 12 mV deviation in LSV curves after 1000 cycles.93 In 2020, Dong et al. employed a simple hydrothermal strategy to synthesize Zn-doped Co9S8 catalysts on copper foam (CF). As shown in Fig. 5h, the morphology of Co9S8@CF without Zn doping exhibited a spherical hierarchical structure; after doping with Zn, the morphology transformed into a neuron-like network. This Zn-doped Co9S8@CF-(1-1) catalyst demonstrates remarkable HER activity in both acidic and alkaline solutions, with low overpotentials (278 mV in 1 M KOH and 273 mV in 0.5 M H2SO4) to reach 10 mA cm−2 and low Tafel slopes (114.4 mV dec−1 in 1 M KOH and 85.2 mV dec−1 in 0.5 M H2SO4), outperforming other samples.95 The comprehensive comparison of the catalytic activities of the in situ grown cobalt sulfides catalysts discussed in this study is shown in Table 3.

Table 3 Comparison of EWS performance of in situ grown cobalt sulfide catalysts reported in the literature
Catalyst Substrate Preparation Overpotential η (mV) at 10 mA cm−2 Tafel slope (mV dec−1) Electrolyte Ref.
HER OER HER OER
Fe0.43Co0.57S2 Metal film Thermal sulfidation ∼190 (4 mA cm−2) 55.9 0.5 M H2SO4 84
CoS/SS Stainless steel Chemical bath deposition 300 57 1 M KOH 88
FeCo2S4/NF Nickel foam Hydrothermal 270 (50 mA cm−2) 59 1 M KOH 94
CoS@NiFe/NF Nickel foam Electrodeposition 175 73.1 1 M KOH 89
Edge-MoS2/Co3S4@NFs Nickel foam Hydrothermal 90.3 61.69 1 M KOH 90
NiCo2S4@NiFe LDH Nickel foam Hydrothermal 200 201 (60 mA cm−2) 101.1 46.3 1 M KOH 92
FeS2-MoS2@CoS2-MOF Nickel foam Thermal sulfidation + hydrothermal 92 211 (20 mA cm−2) 70.4 64.5 1 M KOH 87
FeCo2S4/NF Nickel foam Hydrothermal 132 164 1 M KOH 94
CoS2 NW Graphite Thermal sulfidation 145 51.6 0.5 M H2SO4 85
CoMoS4/SDC Sulfur-doped carbon Hydrothermal 180 48 0.5 M H2SO4 93
Zn-Co9S8@CF-(1-1) Carbon fibre Hydrothermal 273 85.2 0.5 M H2SO4 95
Zn-Co9S8@CF-(1-1) Carbon fibre Hydrothermal 278 114.4 1 M KOH 95
P–Co1−xS Carbon cloth Hydrothermal 110 0.5 M H2SO4 64
Fc-CoxSy/CNT CNT Solvothermal 304 54.2 1 M KOH 91
S, N-CNTs/CoS2@Co N-doped CNT Thermal treatment 112 157 104.9 76.1 1 M KOH 86


3.3 In situ grown nickel sulfides for enhancing EWS

Nickel sulfides have gained significant attention as a versatile class of materials for electrocatalysis. Among these, NiS is notable for its hexagonal structure and good catalytic efficiency, attributed to its ability to facilitate electron transfer and provide active sites for hydrogen adsorption. NiS2, with its distinctive pyrite-like cubic structure, provides a high density of catalytic sites and excellent electrical conductivity, making it particularly effective for the HER. On the other hand, Ni3S2 stands out due to its mixed valence states (+2 and +3) and rhombohedral structure, which contribute to its superior catalytic activity in both the HER and OER by enhancing redox flexibility and stability.
3.3.1 In situ grown NiS and NiS2 electrodes. NiS is recognized for its moderate electrocatalytic activity in the HER. However, according to Jiang's report, NiS exhibited the poorest electrocatalytic activity among nickel sulfides, with an overpotential of 474 mV to reach 10 mA cm−2 current density and a Tafel slope of 124 mV dec−1 in an alkaline medium during the HER process.97 In 2015, Yang et al. improved the catalytic activity of NiS by doping it with tungsten. They synthesized amorphous tungsten-doped nickel sulfide (NiWS) alongside tungsten doped cobalt sulfide (CoWS) for comparison. Unlike crystalline materials, amorphous structures provide more active sites due to their disordered atomic arrangement, potentially leading to higher catalytic activity. The as-prepared NiWS on the FTO substrate demonstrated exceptional HER performance, attributed to the increased number of active sites and improved charge transfer efficiency facilitated by the amorphous structure and composite nature of the material.98 Further enhancement of NiS performance has been achieved through doping strategies. Zhang et al. doped NiS with iron and cobalt, synthesizing a sheet-like FeCoNiSx/NF configuration on nickel foam. This composite structure featured a microporous framework conducive to facilitating oxygen bubble release, which is a critical factor for the continuous OER process. The low crystallization of FeCoNiSx/NF contributed to a high abundance of unsaturated edge sites, enhancing its OER efficiency. Electrochemical assessments confirmed its low Tafel slope of 55 mV dec−1 and a modest overpotential of 231 mV at a current density of 10 mA cm−2, underscoring the composite's efficacy as an OER catalyst.99 Recent advancements have also focused on structure management. Chen et al. reported a one-step solvothermal strategy to prepare a nanoworm-like nickel sulfide structure on NF, which is shown in Fig. 6f (NiS-NW/NF). The as-prepared NiS-NW/NF exhibited significant HER performance with an overpotential of 224 mV at a current density of 20 mA cm−2 and a Tafel slope of 116.24 mV dec−1. Additionally, NiS-NW/NF showed outstanding OER performance, achieving an overpotential of 279 mV at a high current density of 100 mA cm−2, outperforming many recent OER catalysts.105 The good durability of NiS further enhances its potential for catalyzing water splitting. In 2018, Rahman et al. grew NiS on FTO to enhance its electrochemical performance in the HER process. The morphology of the NiS-coated FTO could be controlled by adjusting the annealing temperature. Compared to pristine NiS, NiS-coated FTO demonstrated improved performance, with a lower overpotential of 364 mV at 10 mA cm−2 and a smaller Tafel slope of 90.8 mV dec−1.107
image file: d4ta06197g-f6.tif
Fig. 6 (a) Synthesis diagram of V-incorporated NixSy nanowires; reproduced with permission from ref. 100. Copyright 2017 RSC. (b) HER performance of the as-prepared Co-MOF/CP, Ni3S2/CP, Ni3S2@2D Co-MOF/CP, and 20% Pt/C; reproduced with permission from ref. 101. Copyright 2022 Elsevier. (c) LSV curves of the as-prepared Fe-Ni3S2/FeNi, FeS/Fe, Ni3S2/Ni and IrO2; reproduced with permission from ref. 102. Copyright 2017 Wiley. (d) Preparation schematic of Ni3S2/NF; reproduced with permission from ref. 103. Copyright 2018 RSC. (e) Polarization curves of the as-prepared catalysts, nickel foam and 20 wt% Pt/C/NF for the HER and OER, reproduced with permission from ref. 104. Copyright 2018 ACS. (f) Different morphologies of NiS nanoplates and nanoworms; reproduced with permission from ref. 105. Copyright 2020 Elsevier. (g) LSV results and morphologies of NiS2 catalysts before and after HER activation; reproduced with permission from ref. 106. Copyright 2017 Elsevier.

NiS2, with its pyrite-like cubic structure, outperforms NiS due to its superior catalytic properties, including better electrical conductivity and a greater number of active sites for both the HER and OER. In 2017, Ma et al. employed a combination of hydrothermal synthesis and thermal treatment to prepare nanosheet nickel sulfide on nickel foam (NF-NiS2). The as-prepared NF-NiS2 was further activated after a 5-hour HER process, resulting in an activated sample denoted as NF-NiS2-A. Comparative analysis between pristine NiS2 and NiS2-NF nanosheets revealed enhanced electrochemical activity in the latter, with a lower overpotential of 122 mV at 10 mA cm−2 and a smaller Tafel slope of 86 mV dec−1. Notably, the HER activity of NiS2-NF-A demonstrated a significant increase after the chronopotentiometry test at 20 mA cm−2. The overpotential required to reach 10 mA cm−2 was reduced to 55 mV, and a mere 161 mV overpotential was necessary to reach 100 mA cm−2, closely rivalling the efficiency of the Pt/C catalyst (∼145 mV). Fig. 6g illustrates the morphological changes before and after the HER process, which are attributed to the enhanced catalytic activity.106

3.3.2 In situ grown Ni3S2 electrodes. Ni3S2 typically exhibits the highest catalytic performance among nickel sulfides, owing to its mixed valence states of Ni2+ and Ni3+ combined with a rhombohedral structure that offers a high density of active sites and enhanced redox capabilities. Cathodic activation is an effective method for enhancing the catalytic performance of the HER by modifying the surface properties of catalysts, increasing the availability of active sites and optimizing the electronic structure for hydrogen adsorption. This activation often involves electrochemical treatment that induces structural changes or creates favorable active sites for hydrogen evolution. Shang et al. extensively discussed the in situ cathodic activation (ISCA) of V-incorporated NixSy nanowires for enhanced HER. The ISCA process enhances hydrogen evolution activity by promoting the formation of an amorphous VOOH film on the V-incorporated NixSy nanowires (Fig. 6a), which accelerates the HER process. Additionally, the NiOOH formed during the ISCA process helps protect the active sites, leading to superior activity and structural durability. The activated A-VS/NixSy/NF displayed outstanding HER performance with only 125 mV overpotential required to reach 10 mA cm−2 and a Tafel slope of 113 mV dec−1.100

Depositing Ni3S2 on substrates with excellent conductivity and surface performance is also a good way to enhance its HER performance. In 2022, Cheng et al. employed a solution crystallization method to grow a 2D Co-MOF on treated carbon paper (2D Co-MOF/CP), followed by an in situ electrodeposition strategy to deposit Ni3S2 on the surface (Ni3S2@2D Co-MOF/CP). As shown in Fig. 6b, the as-prepared hierarchical hetero-Ni3S2@2D Co-MOF/CP nanosheets exhibited excellent HER activity with a small overpotential of only 140 mV to reach a current density of 10 mA cm−2 and a Tafel slope of just 90.3 mV dec−1. The 2D structured Co-MOF provided a tunnel for ion transport and sufficient active sites for Ni3S2. This synergy significantly enhanced the HER activity of the as-prepared catalysts, even surpassing that of Pt/C catalysts when the current density exceeded 130 mA cm−2.101

During the OER process, the presence of Ni3+ enhances the material's ability to facilitate the formation and release of oxygen molecules. This redox flexibility is a key factor in lowering the overpotential and increasing the overall efficiency of the OER. In 2017, Yuan et al. explored a simple solvothermal method to grow Ni3S2 nanosheets on FeNi alloy foils. This method simplifies the production of high-efficiency OER electrodes by using FeNi alloy foil as both the Fe and Ni source and the current collector, leading to an in situ grown electrode with a low overpotential (282 mV at a current density of 10 mA cm−2) and high long-term stability in alkaline solutions (Fig. 6c).102 Li et al. utilized metal–organic frameworks (MOFs) as self-sacrificial templates to synthesize oxygen-incorporated Ni3+ self-doped Ni3S4 nanosheets on NF substrates (Ni3S4/NF) in 2020, achieving outstanding catalytic activity. This method resulted in the production of vertically aligned nanosheets with numerous open pores, which are anticipated to enhance the density of active sites for the EWS process. As expected, Ni3S4/NF demonstrated an OER overpotential of 266 mV at 10 mA cm−2 and a Tafel slope of 77 mV dec−1, significantly showing improvement over pristine Ni3S4/NF. The introduction of oxygen into the material matrix notably improved catalytic activity, particularly in the OER, while maintaining stability at higher current densities over extended durations.108

The unique properties of Ni3S2, particularly when synthesized in situ on conductive substrates, contribute to its dual-functionality and high efficiency in both the HER and OER. Ren et al. leveraged the structural benefits of nickel foam as both the substrate and an active component in the catalytic process, facilitating enhanced electrochemical performance through a synergistic effect. Ni3S2/Ni foam is produced through a rapid sulfurization process, where nickel foam is treated to form a nickel sulfide film directly on its surface (Fig. 6d). This method is noteworthy for its simplicity and efficiency, allowing for the creation of a uniform and functional catalytic surface without the need for complex equipment or procedures. The as-prepared Ni3S2/Ni foam exhibited excellent catalytic activity towards water splitting, capable of driving both the HER and OER with smaller overpotentials and larger current densities compared to many conventional catalysts.103 In 2023, Zhu et al. combined electrodeposition and hydrothermal methods to synthesize MoO3–NiSx films with enhanced HER and OER performance by adjusting the deposition duration. The as-prepared samples exhibited bifunctional catalytic performance, with an overpotential of 142 mV for the HER (at 10 mA cm−2) and 294 mV for the OER (at 50 mA cm−2).109 Introducing other elements into catalysts can significantly enhance electrochemical performance and stability by optimizing adsorption energies and modifying the electronic structure.110 In 2018, Jian et al. employed a two-step preparation strategy to synthesize ultrathin Sn-doped Ni3S2 nanosheets on nickel foam. The synthesis method involves a hydrothermal process followed by vulcanization, leading to a catalyst that required notably small overpotentials for both the HER (0.17 V) and OER (0.27 V) to reach 100 mA cm−2 current density. The outstanding catalytic performance is shown in Fig. 6e. The enhanced catalytic activity is attributed to Sn doping, which increases the intrinsic electronic performance of the catalyst. Furthermore, only 1.46 V was required for overall water splitting to achieve a current density of 10 mA cm−2.104 Liu et al. also introduced phosphorus into the iron–nickel sulfide matrix to enhance the electrochemical performance and stability of the catalysts in 2019. The formation of P-(Ni,Fe)3S2 nanosheet arrays directly on an NF substrate provided a large surface area and ensured efficient charge transport. The outstanding catalytic performance of P-(Ni,Fe)3S2/NF for both the HER and OER makes it a versatile electrode for overall water splitting, requiring only 1.54, 1.58, and 1.72 V cell potential to reach 10, 20 and 100 mA cm−2, respectively.111 The comprehensive comparison of the catalytic activities of the in situ grown nickel sulfide catalysts discussed in this study is shown in Table 4.

Table 4 Comparison of EWS performance of in situ grown nickel sulfide catalysts reported in the literature
Catalyst Substrate Preparation Overpotential η (mV) at 10 mA cm−2 Tafel slope (mV dec−1) Electrolyte Ref.
HER OER HER OER
FeCoNiSx/NF Nickel foam Room-temperature sulfuration 231 55 1 M KOH 99
Ni3S4/NF Nickel foam Hydrothermal 266 77 1 M KOH 108
Cu/Ni3S2/NF Nickel foam Hydrothermal 130 84.19 1 M KOH 24
NF-NiS2-A Nickel foam Hydrothermal/thermal sulfurization 67 63 1 M KOH 106
A-VS/NixSy/NF Nickel foam Hydrothermal/electrochemical activation 125 113 1 M KOH 100
NiS-NW/NF Nickel foam Solvothermal 224 (20 mA cm−2) 279 (100 mA cm−2) 116.2 38.44 1 M KOH 105
Ni3S2/Ni foam Nickel foam Mercaptoethanol solution/annealing 131 312 96 111 1 M KOH 103
Sn-Ni3S2/NF Nickel foam Hydrothermal 170 270 55.6 52.7 1 M KOH 104
Fe-Ni3S2/FeNi FeNi alloy foils Solvothermal 282 54 1 M KOH 102
Amorphous NiWS FTO Thermolytic process 250 (8.6 mA cm−2) 55 0.5 M H2SO4 98
NiS/FTO FTO Chemical bath deposition 290 143.4 1 M NaOH 107
MoO3-NiSx (9H5E)/CF Co foam Electrodeposition/hydrothermal 294 (50 mA cm−2) 114 1 M KOH 109
MoO3-NiSx (9H5E)/CC Carbon cloth Electrodeposition/hydrothermal 142 79 1 M KOH 109


3.4 In situ grown TMS composites for enhancing EWS

Having explored the individual contributions of molybdenum, cobalt, and nickel sulfides as active materials in enhancing the catalytic performance of in situ grown electrodes, the focus now shifts to composite TMSs. These composite materials, formed by combining different TMSs, offer a synergistic enhancement of catalytic properties, surpassing the performance of their single-component counterparts. In this section, we will explore various composite TMSs that have been developed to optimize the efficiency and stability of electrocatalytic processes, particularly for the HER and OER. By examining the synthesis methods, structural features, and catalytic mechanisms of these composites, we aim to highlight their potential in pushing the boundaries of water-splitting technologies and other related electrochemical applications.

In 2016, An et al. utilized Ti foil as a substrate to synthesize NiS2–MoS2 catalysts with a hetero-nanowire structure optimized for HER performance. The schematic illustration for preparing NiS2–MoS2 is shown in Fig. 7a. This hybrid structure was created by the thermal conversion of nickel molybdate (NiMoO4) precursors, resulting in a composite with abundant active edge sites and defects. The as-prepared NiS2–MoS2 demonstrates excellent HER activity in basic solutions compared to individual MoS2 or NiS2 nanostructures.112 In 2020, Xue et al. introduced a MoS2/Ni3S2 coaxial heterostructure on nickel foam synthesized via a one-step hydrothermal method followed by annealing. The structure schematic of the as-prepared MoS2/Ni3S2 NW-NF is shown in Fig. 7b. This catalyst exhibits remarkable HER performance at high current density, requiring overpotentials of 182 and 200 mV to reach current densities of 500 and 1000 mA cm−2, respectively.113 The OER performance of these materials has also been investigated. Dong and colleagues synthesized a NiFeS/NF catalyst through a two-step method. Initially, nanosheets of NiFe hydroxide were electrodeposited onto NF, creating a precursor with an extensive surface area. This was followed by a hydrothermal sulfurization process, which led to the formation of NiFeS/NF, an effective electrocatalyst for the OER. This synthesis approach enhanced the material's catalytic properties by incorporating sulfur into the NiFe structure, thereby improving its activity and stability during the oxygen evolution reaction. XRD results revealed mixed phases of NiS and Ni3S2. As shown in Fig. 7c, the outstanding OER performance was characterized by a small overpotential of 65 mV at a current density of 10 mA cm−2, coupled with a large electrochemical active surface area of 251.25 cm−2.114


image file: d4ta06197g-f7.tif
Fig. 7 (a) Schematic illustration for preparing NiS2-MoS2 using sulfurization and CVD strategies; reproduced with permission from ref. 112. Copyright 2016 RSC. (b) Structure schematic of the as-prepared MoS2/Ni3S2 NW-NF; reproduced with permission from ref. 113. Copyright Elsevier 2020. (c) LSV curves and double-layer capacitance of the as-prepared catalysts for the HER; reproduced with permission from ref. 114. Copyright 2016 RSC. (d) Schematic illustration of the process of overall water splitting with FeCoNiP0S1 and FeCoNiP0.5S0.5 as the electrodes along with LSV curves of the as-prepared catalysts for the HER, reproduced with permission from ref. 115. Copyright 2018 ACS. (e) Diagram and polarization curves of NiFeS/NF-P for the OER; reproduced with permission from ref. 116. Copyright 2018 Elsevier. (f) LSV curves of the as-prepared catalysts for the HER; reproduced with permission from ref. 117. Copyright 2020 Elsevier. (g) Schematic illustration for the preparation of CoMoSx on nickel foam with its unique structure; reproduced with permission from ref. 118. Copyright 2021 Elsevier.

The ultimate goal in electrocatalysis is to develop materials that can efficiently catalyse both the HER and OER within the same system. Wang et al. synthesized amorphous multi-element electrocatalysts on a Ti sheet in 2018, integrating Fe, Co, Ni, S, P, and O through a singular electrodeposition strategy for water splitting. The as-prepared catalysts, named FeCoNiPxSy, were designed to modulate the non-metal composition, thereby altering their electrochemical performance for both the HER and OER. The schematic diagram is shown in Fig. 7d. The catalytic activity results indicate that S and P play crucial roles in enhancing the catalysts' intrinsic and apparent activities, highlighting a pathway toward optimizing multi-element catalysts for efficient electrocatalysis.115 In the same year, Yan et al. employed an electrodeposition strategy to deposit ternary NiFeMoS anemone-like nanorods on nickel foam. They investigated the influence of pluronic P123 and MoO42− on the nanostructure and electrocatalytic activity of NiFeMoS. The anemone-like nanorods were formed due to the presence of P123 and MoO42−, which resulted in a scaly surface on NiFeMoS. The NiFeMoS/NF-P catalyst exhibits outstanding electrocatalytic performance, achieving a significant current density of 150 mA cm−2 at a relatively low overpotential for the OER (280 mV) in alkaline media. Moreover, the HER performance of this as-prepared catalyst was also excellent, with a 100 mV overpotential at a current density of 10 mA cm−2. Consequently, the full cell assembled with the as-prepared NiFeS/NF-P catalyst displayed a very low water electrolysis voltage of 1.52 V to reach 100 mA cm−2 (Fig. 7e).116 In 2019, Shit et al. successfully synthesized a unique CoSx/Ni3S2 heterostructure, which was skillfully integrated with nickel foam (CoSx/Ni3S2@NF) using a one-pot hydrothermal method. Unlike many catalysts that require complex preparation or additional support materials, in situ grown CoSx/Ni3S2@NF ensured strong adhesion and excellent electrical connectivity. The as-prepared CoSx/Ni3S2@NF required remarkably low overpotentials (204 and 280 mV) to achieve significant current densities (10 and 20 mA cm−2) for the HER and OER.119 In 2020, Huang et al. reported a two-step synthesis method involving the hydrothermal growth of NiMo precursor cuboids on nickel foam, followed by nitrogen doping via vulcanization. The as-prepared N-NiMoS/NF exhibited a low overpotential of 68 mV at 10 mA cm−2 current density and an 86 mV dec−1 Tafel slope for the HER, demonstrating performance comparable to that of the most advanced Pt/C catalysts. Furthermore, the as-prepared catalyst maintained outstanding HER performance at high current density (Fig. 7f).117 Zhao et al. presented an innovative approach for fabricating a hierarchical electrocatalyst composed of Ni3S2-CoMoSx on NF in 2020. As shown in Fig. 7g, by employing nickel foam as a conductive substrate, the study leveraged its 3D porous structure to facilitate electron transport and provide a large surface area for the uniform growth of the catalytic materials. The electrocatalyst demonstrated exceptional bifunctional activity for overall water splitting.118

In the previous section, we thoroughly explored the roles of molybdenum, cobalt, and nickel sulfides, as well as their composites, as active materials in enhancing the catalytic performance of in situ grown electrodes. These materials have been shown to significantly boost the efficiency and stability of electrochemical reactions, particularly in the context of the HER and OER. By combining different TMSs, we have seen how their composite form can achieve even greater catalytic performance, leveraging the synergistic effects between various metal elements. The comprehensive comparison of the catalytic activities of the in situ grown composite TMS catalysts discussed in this study is shown in Table 5.

Table 5 Comparison of EWS performance of in situ grown composite TMS catalysts reported in the literature
Catalyst Substrate Preparation Overpotential η (mV) at 10 mA cm−2 Tafel slope (mV dec−1) Electrolyte Ref.
HER OER HER OER
NiS2-MoS2/Ti foil Ti foil Solvothermal + chemical vapor deposition 200 (onset potential) 70 1 M KOH 112
FeCoNiP0S1 Ti foil Electrodeposition 135 (100 mA cm−2) 99 1 M KOH 115
FeCoNiP0.5S0.5 Ti foil Electrodeposition 258 (100 mA cm−2) 49 1 M KOH 115
FS-0.9-30 Nickel foam Electrodeposition 157 112.5 1 M KOH 120
N-NiMoS/NF Nickel foam Hydrothermal 68 86 1 M KOH 117
Ni3S2-CoMoSx/NF Nickel foam Sulfidation 234 90 125 75 1 M KOH 118
P-(Ni,Fe)3S2/NF Nickel foam Phosphorization + sulfuration 98 196 88 30 1 M KOH 111
NiFeMoS/NF-P Nickel foam Electrodeposited + hydrothermal sulfuration 100 280 (150 mA cm−2) 121 69 1 M KOH 116
Co9S8-Ni3S2 HNTs/Ni Nickel foam Hydrothermal 85 281 (50 mA cm−2) 83.1 53.3 1 M KOH 121
NF-Ni3S2/MnO2 Nickel foam Hydrothermal 102 260 69 61 1 M KOH 122
(Ni-Fe)Sx/NiFe(OH)y/NF Nickel foam Electrodeposition 124 (100 mA cm−2) 290 (100 mA cm−2) 97 58 1 M KOH 123
Ni–S–B/NM Nickel mesh Electrodeposition 240 121.2 30 wt% KOH 124
Co:WS2/Co:W18O4/FTO FTO Thermal sulfidation 210 49 0.5 M H2SO4 125
1T/2H-MoS2@CC Carbon cloth Hydrothermal 114 52.8 0.5 M H2SO4 126
NS-500 N/S co-doped graphene Chemical vapor deposition 130 (onset potential) 80.5 0.5 M H2SO4 127
MoSx/CNT-rGO/VTP CNT-rGO Hydrothermal 190 (onset potential) 59 0.5 M H2SO4 128
Co(II)1−xCo(0)x/3Mn(III)2x/3S B/N-codoped mesoporous nanocarbon In situ pyrolysis + hydrothermal 260 (20 mA cm−2) 50 1 M KOH 129


The exploration of TMSs, particularly in the context of in situ grown catalysts, has demonstrated significant potential in advancing the field of electrocatalysis. Through the comprehensive analysis of molybdenum, cobalt, and nickel sulfides, this study highlights the substantial improvements in catalytic performance, particularly for the HER and OER, when these materials are synthesized directly on conductive substrates. The superior electrical conductivity, increased density of active sites, and enhanced redox flexibility offered by these TMSs underscore their potential as efficient, cost-effective alternatives to traditional noble metal catalysts.

4. Conclusion and prospects

In conclusion, the pursuit of sustainable and efficient energy solutions has underscored the pivotal importance of electrocatalytic water splitting (EWS) as a transformative technology for hydrogen production, a clean and renewable energy source. The in situ grown TMSs on conductive substrates have been identified as a compelling alternative to noble metal-based catalysts due to their cost-effectiveness, structural diversity, and abundant active sites, which not only enhance electron transfer and catalytic performance but also eliminate the need for polymer binders, further improving electrode stability and reducing preparation costs. In this work, we first reviewed the three basic elements of EWS using in situ grown TMS-based catalysts, i.e., the TMS materials, suitable substrates for the in situ growth process, and the in situ growth methods.

Because the TMS materials are directly grown on the substrate, the performance characteristics of the substrate material crucially affect the performance of the synthesized catalysts. Therefore, the selection of substrate materials for in situ grown catalysts is crucial for achieving high catalytic efficiency in applications. Herein, carbon-based and nickel foam substrates are discussed in detail. Carbon-based materials, including carbon nanotubes, graphene, and carbon cloth, are characterized by their exceptional electrical conductivity, extensive surface area, and strong mechanical stability. These properties make them ideal substrates for supporting active catalytic materials. Nickel foam is a porous metallic substrate known for its high electrical conductivity, mechanical strength, and three-dimensional (3D) structure. The unique 3D porous network of nickel foam offers a large active surface area and facilitates mass transport, which is essential for efficient catalysis.

Furthermore, the most commonly adopted in situ growth techniques, hydrothermal synthesis and electrodeposition, are comprehensively investigated, offering the following advantages: (1) Direct growth on conductive substrates: they enable the in situ growth of active materials on substrates, facilitating the fabrication of cohesive, binder-free electrodes. This direct integration enhances electron transport and reduces interfacial resistance, beneficial for electrochemical applications. (2) Controlled morphology and structure: both these techniques allow for meticulous control over the morphology and structure of the resulting materials. In hydrothermal synthesis, parameters such as solvent composition, temperature, and pressure can be adjusted to tailor the size, shape, and arrangement of nanostructures. (3) Versatility in material types and compositions: both methods support the synthesis of a wide range of materials, including sulfides, metals, metal oxides, and hybrid composites. This versatility allows for the exploration of novel material systems and compositions tailored for specific applications. (4) Energy and cost efficiency: compared to other synthetic techniques, hydrothermal and electrodeposition processes are relatively energy-efficient and cost-effective, making them suitable for large-scale production.

While in situ grown TMSs have demonstrated promising catalytic performance as electrocatalysts for water splitting, several challenges and developmental shortcomings remain to be addressed:

(1) The durability and long-term operational stability of in situ grown TMS catalysts, which are crucial for their practical application in electrocatalytic processes, can be significantly affected by various factors. One such factor is the dissolution of catalytic materials, which leads to a gradual loss of active sites and a consequent decrease in catalytic activity. Additionally, structural degradation characterized by morphological changes, surface corrosion, or phase transformation under harsh operational conditions, can adversely affect the physical integrity and electrochemical performance of the catalysts. Together, these factors pose significant challenges to the efficiency and longevity of in situ grown TMS catalysts in efficient water splitting.

(2) Maximizing the exposure of catalytically active sites and increasing the surface area of in situ grown TMS catalysts are crucial for improving their performance. Advanced synthesis techniques that can control the nanostructure and morphology of TMS catalysts on the surfaces of substrates are needed. Such techniques should enable the precise manipulation of catalyst features including particle size, shape, and distribution, as well as the engineering of porosity at the nanoscale, directly on the surface of conductive substrates. By fine-tuning these parameters, it is possible to significantly enhance the active surface area, thereby facilitating a higher rate of catalytic reactions. Moreover, the introduction of structural features such as hierarchical porosities and interconnected networks can further improve mass transfer and gas diffusion efficiency, optimizing the overall catalytic performance.

(3) The interface between the TMS catalyst and the substrate plays a critical role in the overall catalytic performance. Optimizing this interface through better material integration and interfacial engineering can enhance the efficiency and durability of the catalysts. Better integration of TMS catalysts onto substrates involves creating a seamless connection that facilitates efficient electron transfer and minimizes interfacial resistance. Interfacial engineering, on the other hand, focuses on modifying the surface properties at the interface to improve catalytic performance. This can include the introduction of functional groups, dopants, or co-catalysts that can modulate the electronic structure of the interface, thereby optimizing the adsorption and desorption energies of reactants and intermediates.

(4) Some other TMSs with potential catalytic capabilities have not been covered in detail due to their relatively unexplored nature. Although these materials may not yet have the extensive research backing of the more established sulfides, they represent promising avenues for future exploration and development in the field of electrocatalysis. For example, Shi et al. adopted a multi-step strategy involving flame vapor deposition, sol-flame doping, and sulfurization to prepare Co-doped WS2 with excellent HER activity.125 Continued research into these lesser-known TMSs could uncover new insights and further expand the toolkit available for designing high-performance catalytic materials.

(5) Electrodeposition, through the manipulation of deposition parameters like voltage, current density, and electrolyte composition, offers precise control over the thickness, composition, and microstructure of the deposited layers. However, this technique also presents challenges, such as ensuring uniform coverage over complex geometries. Additionally, the choice of electrolyte and deposition conditions can significantly influence the crystallinity and phase purity of the resulting material, which in turn affects its electrocatalytic performance.

Addressing the challenges associated with in situ grown TMS catalysts will require a multi-faceted approach that includes improving structural stability, optimizing synthesis techniques, enhancing interfacial properties, and expanding the exploration of lesser-known materials. In the future, doping and heterostructure formation will play pivotal roles in overcoming conductivity limitations and improving the kinetics of catalytic reactions. These strategies can significantly enhance the efficiency, durability, and practical applications of TMS materials in electrocatalysis.

In conclusion, in situ grown TMS catalysts hold great promise for advancing electrochemical energy conversion technologies. In the future, more efforts should be focused on the development of novel and scalable synthesis techniques that allow for precise control over the morphology, crystallinity, and composition of TMS catalysts. Meanwhile, the creation of hybrid materials by combining TMSs with conductive carbon materials, polymers, or other metal oxides offers great potential for improving both conductivity and stability. Continued exploration of lesser-known TMS materials with novel compositions, such as mixed-metal sulfides and high-entropy TMSs, may lead to the discovery of new active materials with superior catalytic properties. We hope this review can provide a comprehensive overview of the latest progress and remaining challenges, serving as a critical resource for researchers and engineers seeking to optimize these materials further.

Data availability

No primary research results, software or code have been included and no new data were generated or analysed as part of this review.

Author contributions

H. W. Jia, P. Y. Guan, L. Zhou, Y. Z. Zhou and D. W. Chu devised and outlined the draft of the review paper. H. W. Jia contributed to the scientific writing of the manuscript. H. W. Jia, L. H. Meng, Z. K. Dong, C. Liu, and P. Y. Guan constructed the figures for illustrations. H. W. Jia, L. H. Meng, Y. L. Lu, T. Y. Liang, Y. Yu and Y. F. Hu constructed the tables for illustrations. H. W. Jia, Y. Z. Zhou, P. Y. Guan, M. Y. Li, T. Wan, B. J. Ni, Z. J. Han, and D. W. Chu contributed to the review and editing of the manuscript. All authors contributed to the final polishing of the manuscript.

Conflicts of interest

The authors declare that they have no conflict of interest.

Acknowledgements

Haowei Jia would like to acknowledge the financial support from the University International Postgraduate Award (UIPA) funded by the University of New South Wales, Australia.

References

  1. L. Lu, et al., Ligand-free synthesis of noble metal nanocatalysts for electrocatalysis, Chem. Eng. J., 2023, 451, 138668 CrossRef CAS .
  2. T. Zhang, et al., Mo2B2 MBene-supported single-atom catalysts as bifunctional HER/OER and OER/ORR electrocatalysts, J. Mater. Chem. A, 2021, 9(1), 433–441 RSC .
  3. H. Begum, M. S. Ahmed and S. Jeon, δ-MnO2 nanoflowers on sulfonated graphene sheets for stable oxygen reduction and hydrogen evolution reaction, Electrochim. Acta, 2019, 296, 235–242 CrossRef CAS .
  4. A. S. Agnihotri, A. Varghese and N. M, Transition metal oxides in electrochemical and bio sensing: A state-of-art review, Appl. Surf. Sci. Adv., 2021, 4, 100072 CrossRef .
  5. H. Jia, et al., Nickel-Doped Manganese Dioxide Electrocatalysts with MXene Surface Decoration for Oxygen Evolution Reaction, Energy Fuels, 2022, 36(22), 13808–13816 CrossRef CAS .
  6. Y. Zhou, et al., Engineering work functions of cobalt-doped manganese oxide based electrocatalysts for highly efficient oxygen evolution reaction, J. Colloid Interface Sci., 2023, 642, 23–28 CrossRef CAS PubMed .
  7. Y.-N. Zhou, et al., Carbon-based transition metal sulfides/selenides nanostructures for electrocatalytic water splitting, J. Alloys Compd., 2021, 852, 156810 CrossRef CAS .
  8. J. Liang, et al., Multiphase interface coupling of Ni-based sulfide composites for high-current-density oxygen evolution electrocatalysis in alkaline freshwater/simulated seawater/seawater, Dalton Trans., 2024, 53, 15040–15047 RSC .
  9. Y. Li, et al., Interface engineering of transitional metal sulfide–MoS2 heterostructure composites as effective electrocatalysts for water-splitting, J. Mater. Chem. A, 2021, 9(4), 2070–2092 RSC .
  10. R. Li, et al., Electrodeposition: Synthesis of advanced transition metal-based catalyst for hydrogen production via electrolysis of water, J. Energy Chem., 2021, 57, 547–566 CrossRef CAS .
  11. X. Yu, et al., Highly disordered cobalt oxide nanostructure induced by sulfur incorporation for efficient overall water splitting, Nano Energy, 2020, 71, 104652 CrossRef .
  12. X. Zheng, et al., Rational Design and Spontaneous Sulfurization of NiCo-(oxy)Hydroxysulfides Nanosheets with Modulated Local Electronic Configuration for Enhancing Oxygen Electrocatalysis, Adv. Energy Mater., 2022, 12(15), 2103275 CrossRef .
  13. M. Cui, et al., High-Entropy Metal Sulfide Nanoparticles Promise High-Performance Oxygen Evolution Reaction, Adv. Energy Mater., 2021, 11(3), 2002887 CrossRef .
  14. C. Liu, et al., Unveil the Triple Roles of Water Molecule on Power Generation of MXene Derived TiO2 based Moisture Electric Generator, Adv. Energy Mater., 2024, 14(27), 2400590 CrossRef .
  15. H. Tong, et al., Zinc cobalt sulfide nanosheets grown on nitrogen-doped graphene/carbon nanotube film as a high-performance electrode for supercapacitors, J. Mater. Chem. A, 2016, 4(29), 11256–11263 RSC .
  16. P. Kulkarni, et al., Nanostructured binary and ternary metal sulfides: synthesis methods and their application in energy conversion and storage devices, J. Mater. Chem. A, 2017, 5(42), 22040–22094 RSC .
  17. J.-M. Kim, et al., A review on the stability and surface modification of layered transition-metal oxide cathodes, Mater. Today, 2021, 46, 155–182 CrossRef .
  18. C. Maheu, et al., Titania – Supported transition metals sulfides as photocatalysts for hydrogen production from propan-2-ol and methanol, Int. J. Hydrogen Energy, 2019, 44(33), 18038–18049 CrossRef .
  19. Y. Yang, et al., Recent advances in application of transition metal phosphides for photocatalytic hydrogen production, Chem. Eng. J., 2021, 405, 126547 CrossRef .
  20. Y. Li, et al., MoS2 Nanoparticles Grown on Graphene: An Advanced Catalyst for the Hydrogen Evolution Reaction, J. Am. Chem. Soc., 2011, 133(19), 7296–7299 CrossRef .
  21. J. Yang, et al., Two-Dimensional Hybrid Nanosheets of Tungsten Disulfide and Reduced Graphene Oxide as Catalysts for Enhanced Hydrogen Evolution, Angew. Chem., Int. Ed., 2013, 52(51), 13751–13754 CrossRef .
  22. S. Chang, et al., In situ Growth of Silver Nanoparticles in Porous Membranes for Surface-Enhanced Raman Scattering, ACS Appl. Mater. Interfaces, 2010, 2(11), 3333–3339 CrossRef .
  23. W. Zeng, et al., In situ synthesis of binded, thick and porous carbon nanoparticle dye sensitized solar cell counter electrode with nickel gel as catalyst source, J. Power Sources, 2014, 245, 456–462 CrossRef .
  24. Q. Xiao, et al., Deposition of Cu on Ni3S2 nanomembranes with simply spontaneous replacement reaction for enhanced hydrogen evolution reaction, J. Electroanal. Chem., 2022, 911, 116214 CrossRef .
  25. T. Zhang, J. Sun and J. Guan, Self-supported transition metal chalcogenides for oxygen evolution, Nano Res., 2023, 16(7), 8684–8711 CrossRef .
  26. J. Liu, et al., Self-Supported Earth-Abundant Nanoarrays as Efficient and Robust Electrocatalysts for Energy-Related Reactions, ACS Catal., 2018, 8(7), 6707–6732 CrossRef .
  27. B. L. Ellis, P. Knauth and T. Djenizian, Three-Dimensional Self-Supported Metal Oxides for Advanced Energy Storage, Adv. Mater., 2014, 26(21), 3368–3397 CrossRef .
  28. R. G. Dickinson and L. Pauling, The crystal structure of molybdenite, J. Am. Chem. Soc., 1923, 45(6), 1466–1471 CrossRef .
  29. J. A. Wilson and A. D. Yoffe, The transition metal dichalcogenides discussion and interpretation of the observed optical, electrical and structural properties, Adv. Phys., 1969, 18(73), 193–335 CrossRef .
  30. Y. Guo, Nanoarchitectonics for Transition-Metal-Sulfide-Based Electrocatalysts for Water Splitting, Adv. Mater., 2019, 31(17), 1807134 CrossRef .
  31. Y. Zhao, et al., Recent advances in defect-engineered molybdenum sulfides for catalytic applications, Mater. Horiz., 2023, 10(10), 3948–3999 RSC .
  32. K. Weiss and J. M. Phillips, Calculated specific surface energy of molybdenite (MoS2), Phys. Rev. B: Condens. Matter Mater. Phys., 1976, 14(12), 5392–5395 CrossRef .
  33. S. Das, G. Swain and K. Parida, One step towards the 1T/2H-MoS2 mixed phase: a journey from synthesis to application, Mater. Chem. Front., 2021, 5(5), 2143–2172 RSC .
  34. X. Chia, et al., Electrochemistry of Nanostructured Layered Transition-Metal Dichalcogenides, Chem. Rev., 2015, 115(21), 11941–11966 CrossRef PubMed .
  35. L. Li and A. Ghahreman, Hydrothermal Monodisperse Microspherulite Pyrite: Novel Synthesis Process and Electrochemical Study of Its Oxidation, ACS Omega, 2020, 5(38), 24871–24880 CrossRef PubMed .
  36. S. Anantharaj, S. Kundu and S. Noda, Progress in nickel chalcogenide electrocatalyzed hydrogen evolution reaction, J. Mater. Chem. A, 2020, 8(8), 4174–4192 RSC .
  37. C. V. V. M. Gopi, Facile Synthesis of Battery-Type CuMn2O4 Nanosheet Arrays on Ni Foam as an Efficient Binder-Free Electrode Material for High-Rate Supercapacitors, Nanomaterials, 2023, 13(6), 1125 CrossRef .
  38. B. Chen, et al., Preparation of MoS2/TiO2 based nanocomposites for photocatalysis and rechargeable batteries: progress, challenges, and perspective, Nanoscale, 2018, 10(1), 34–68 RSC .
  39. M.-R. Gao, et al., Pyrite-Type Nanomaterials for Advanced Electrocatalysis, Acc. Chem. Res., 2017, 50(9), 2194–2204 CrossRef .
  40. F. Yan, et al., Electrochemically activated-iron oxide nanosheet arrays on carbon fiber cloth as a three-dimensional self-supported electrode for efficient water oxidation, J. Mater. Chem. A, 2016, 4(16), 6048–6055 RSC .
  41. X. Han, Ultrafine Pt Nanoparticle-Decorated Pyrite-Type CoS2 Nanosheet Arrays Coated on Carbon Cloth as a Bifunctional Electrode for Overall Water Splitting, Adv. Energy Mater., 2018, 8(24), 1800935 CrossRef .
  42. M.-M. Wang, et al., Multimetallic CuCoNi Oxide Nanowires In Situ Grown on a Nickel Foam Substrate Catalyze Persulfate Activation via Mediating Electron Transfer, Environ. Sci. Technol., 2022, 56(17), 12613–12624 CrossRef CAS .
  43. H. Tian, et al., Self-supported CoMoO4/NiFe-LDH core–shell nanorods grown on nickel foam for enhanced electrocatalysis of oxygen evolution, Dalton Trans., 2022, 51(36), 13762–13770 RSC .
  44. H. Qian, et al., A ternary hybrid of Zn-doped MoS2-RGO for highly effective electrocatalytic hydrogen evolution, J. Colloid Interface Sci., 2021, 599, 100–108 CrossRef CAS PubMed .
  45. D. Huang, et al., The synergistic effect of proton intercalation and electron transfer via electro-activated molybdenum disulfide/graphite felt toward hydrogen evolution reaction, J. Catal., 2020, 381, 175–185 CrossRef CAS .
  46. K. Li, et al., Carbon-Based Fibers: Fabrication, Characterization and Application, Adv. Fiber Mater., 2022, 4(4), 631–682 CrossRef CAS .
  47. Y. Lu, et al., Mechanisms elucidation of secondary seawater batteries: From ion migration to conversion for sustainable energy storage, Chem. Eng. J., 2024, 498, 155307 CrossRef CAS .
  48. P. Geng, Transition Metal Sulfides Based on Graphene for Electrochemical Energy Storage, Adv. Energy Mater., 2018, 8(15), 1703259 CrossRef .
  49. X. Xiong, et al., Three-dimensional ultrathin Ni(OH)2 nanosheets grown on nickel foam for high-performance supercapacitors, Nano Energy, 2015, 11, 154–161 CrossRef CAS .
  50. N. K. Chaudhari, et al., Nanostructured materials on 3D nickel foam as electrocatalysts for water splitting, Nanoscale, 2017, 9(34), 12231–12247 RSC .
  51. Y. Qiang, H. Li and X. Lan, Self-assembling anchored film basing on two tetrazole derivatives for application to protect copper in sulfuric acid environment, J. Mater. Sci. Technol., 2020, 52, 63–71 CrossRef CAS .
  52. S. Zhang, et al., Laser surface alloying of FeCoCrAlNi high-entropy alloy on 304 stainless steel to enhance corrosion and cavitation erosion resistance, Opt Laser. Technol., 2016, 84, 23–31 CrossRef CAS .
  53. D. V. Girenko, et al., Electrodeposition of thin electrocatalytic PbO2 layer on fluorine-doped tin oxide substrates, J. Electroanal. Chem., 2014, 712, 194–201 CrossRef CAS .
  54. D.-J. Guo and Z.-H. Jing, Electrocatalytic properties of platinum nanoparticles supported on fluorine tin dioxide/multi-walled carbon nanotube composites for methanol electrooxidation in acidic medium, J. Colloid Interface Sci., 2011, 359(1), 257–260 CrossRef CAS .
  55. R. Bel Hadj Tahar, et al., Tin doped indium oxide thin films: Electrical properties, J. Appl. Phys., 1998, 83(5), 2631–2645 CrossRef CAS .
  56. M. Z. Alam, I. De Leon and R. W. Boyd, Large optical nonlinearity of indium tin oxide in its epsilon-near-zero region, Science, 2016, 352(6287), 795–797 CrossRef CAS PubMed .
  57. H. M. L. Robert, et al., Light-Assisted Solvothermal Chemistry Using Plasmonic Nanoparticles, ACS Omega, 2016, 1(1), 2–8 CrossRef CAS .
  58. M. Xie, et al., In situ hydrothermal deposition as an efficient catalyst supporting method towards low-temperature graphitization of amorphous carbon, Carbon, 2014, 77, 215–225 CrossRef CAS .
  59. I. L. Peczak, et al., Scalable Synthesis of Pt/SrTiO3 Hydrogenolysis Catalysts in Pursuit of Manufacturing-Relevant Waste Plastic Solutions, ACS Appl. Mater. Interfaces, 2021, 13(49), 58691–58700 CrossRef CAS PubMed .
  60. Z. Ma, et al., Three-dimensional well-mixed/highly-densed NiS-CoS nanorod arrays: An efficient and stable bifunctional electrocatalyst for hydrogen and oxygen evolution reactions, Electrochim. Acta, 2018, 260, 82–91 CrossRef CAS .
  61. H. Che, et al., Hydrothermal electrochemical deposition synthesis NiSe2 as efficient counter electrode materials for dye-sensitized solar cells, J. Alloys Compd., 2017, 705, 645–651 CrossRef CAS .
  62. J. Lu, J. W. Elam and P. C. Stair, Synthesis and Stabilization of Supported Metal Catalysts by Atomic Layer Deposition, Acc. Chem. Res., 2013, 46(8), 1806–1815 CrossRef CAS .
  63. M. Wang, et al., Recent advances in transition-metal-sulfide-based bifunctional electrocatalysts for overall water splitting, J. Mater. Chem. A, 2021, 9(9), 5320–5363 RSC .
  64. G. Qu, et al., Rational design of phosphorus-doped cobalt sulfides electrocatalysts for hydrogen evolution, Nano Res., 2019, 12(12), 2960–2965 CrossRef CAS .
  65. K. Tekin, S. Karagöz and S. Bektaş, A review of hydrothermal biomass processing, Renew. Sustain. Energy Rev., 2014, 40, 673–687 CrossRef .
  66. Ü. Ağbulut, et al., Microalgae bio-oil production by pyrolysis and hydrothermal liquefaction: Mechanism and characteristics, Bioresour. Technol., 2023, 376, 128860 CrossRef PubMed .
  67. F. Guo, Recent Advances in Ultralow-Pt-Loading Electrocatalysts for the Efficient Hydrogen Evolution, Adv. Sci., 2023, 10(21), 2301098 CrossRef PubMed .
  68. J. R. Zeng, et al., Facile electrodeposition of cauliflower-like S-doped nickel microsphere films as highly active catalysts for electrochemical hydrogen evolution, J. Mater. Chem. A, 2017, 5(29), 15056–15064 RSC .
  69. S. Chen, T. Takata and K. Domen, Particulate photocatalysts for overall water splitting, Nat. Rev. Mater., 2017, 2(10), 17050 CrossRef .
  70. Y. Bai, et al., Engineering the surface charge states of nanostructures for enhanced catalytic performance, Mater. Chem. Front., 2017, 1(10), 1951–1964 RSC .
  71. J. Cao, et al., Super-assembled carbon nanofibers decorated with dual catalytically active sites as bifunctional oxygen catalysts for rechargeable Zn-air batteries, Mater. Today Energy, 2021, 20, 100682 CrossRef CAS .
  72. V. Chakrapani, Probing Active Sites and Reaction Intermediates of Electrocatalysis Through Confocal Near-Infrared Photoluminescence Spectroscopy: A Perspective, Front. Chem., 2020, 8, 00327 CrossRef CAS .
  73. X. Zhang, et al., A MnCo2S4 nanowire array as an earth-abundant electrocatalyst for an efficient oxygen evolution reaction under alkaline conditions, J. Mater. Chem. A, 2017, 5(33), 17211–17215 RSC .
  74. J. Joyner, et al., Graphene Supported MoS2 Structures with High Defect Density for an Efficient HER Electrocatalysts, ACS Appl. Mater. Interfaces, 2020, 12(11), 12629–12638 CrossRef CAS .
  75. W. Dong, et al., Defective-MoS2/rGO heterostructures with conductive 1T phase MoS2 for efficient hydrogen evolution reaction, Int. J. Hydrogen Energy, 2021, 46(14), 9360–9370 CrossRef CAS .
  76. C. Zhao, et al., Construction of amorphous CoFeOx(OH)y/MoS2/CP electrode for superior OER performance, Int. J. Hydrogen Energy, 2022, 47(67), 28859–28868 CrossRef CAS .
  77. Y.-J. Tang, et al., Molybdenum Disulfide/Nitrogen-Doped Reduced Graphene Oxide Nanocomposite with Enlarged Interlayer Spacing for Electrocatalytic Hydrogen Evolution, Adv. Energy Mater., 2016, 6(12), 1600116 CrossRef .
  78. P. Xiong, et al., Unilamellar Metallic MoS2/Graphene Superlattice for Efficient Sodium Storage and Hydrogen Evolution, ACS Energy Lett., 2018, 3(4), 997–1005 CrossRef CAS .
  79. A.-Y. Lu, High-Sulfur-Vacancy Amorphous Molybdenum Sulfide as a High Current Electrocatalyst in Hydrogen Evolution, Small, 2016, 12(40), 5530–5537 CrossRef CAS .
  80. Y. Cheng, Rh MoS2 Nanocomposite Catalysts with Pt-Like Activity for Hydrogen Evolution Reaction, Adv. Funct. Mater., 2017, 27(23), 1700359 CrossRef .
  81. M. Li, et al., Vanadium doped 1T MoS2 nanosheets for highly efficient electrocatalytic hydrogen evolution in both acidic and alkaline solutions, Chem. Eng. J., 2021, 409, 128158 CrossRef CAS .
  82. T. Liang, et al., A facile approach to enhance the hydrogen evolution reaction of electrodeposited MoS2 in acidic solutions, New J. Chem., 2022, 46(48), 23344–23350 RSC .
  83. Y. Yin, et al., Contributions of Phase, Sulfur Vacancies, and Edges to the Hydrogen Evolution Reaction Catalytic Activity of Porous Molybdenum Disulfide Nanosheets, J. Am. Chem. Soc., 2016, 138(25), 7965–7972 CrossRef CAS .
  84. D. Kong, et al., First-row transition metal dichalcogenide catalysts for hydrogen evolution reaction, Energy Environ. Sci., 2013, 6(12), 3553–3558 RSC .
  85. M. S. Faber, et al., High-Performance Electrocatalysis Using Metallic Cobalt Pyrite (CoS2) Micro- and Nanostructures, J. Am. Chem. Soc., 2014, 136(28), 10053–10061 CrossRef CAS .
  86. J.-Y. Wang, et al., S, N co-doped carbon nanotube-encapsulated core-shelled CoS2@Co nanoparticles: efficient and stable bifunctional catalysts for overall water splitting, Sci. Bull., 2018, 63(17), 1130–1140 CrossRef CAS PubMed .
  87. K. Chhetri, et al., Engineering the abundant heterointerfaces of integrated bimetallic sulfide-coupled 2D MOF-derived mesoporous CoS2 nanoarray hybrids for electrocatalytic water splitting, Mater. Today Nano, 2022, 17, 100146 CrossRef CAS .
  88. S. B. Kale, et al., Cobalt sulfide thin films for electrocatalytic oxygen evolution reaction and supercapacitor applications, J. Colloid Interface Sci., 2018, 532, 491–499 CrossRef CAS PubMed .
  89. M. Wang, Synergistic Assembly of a CoS@NiFe/Ni Foam Heterostructure Electrocatalyst for Efficient Water Oxidation Catalysis at Large Current Densities, Chem. - Asian J., 2020, 15(9), 1484–1492 CrossRef CAS PubMed .
  90. O. Peng, et al., Hierarchical heterostructured nickle foam–supported Co3S4 nanorod arrays embellished with edge-exposed MoS2 nanoflakes for enhanced alkaline hydrogen evolution reaction, Mater. Today Energy, 2020, 18, 100513 CrossRef CAS .
  91. P. Thangasamy, Ferrocene-Incorporated Cobalt Sulfide Nanoarchitecture for Superior Oxygen Evolution Reaction, Small, 2020, 16(31), 2001665 CrossRef .
  92. J. Liu, et al., Hierarchical NiCo2S4@NiFe LDH Heterostructures Supported on Nickel Foam for Enhanced Overall-Water-Splitting Activity, ACS Appl. Mater. Interfaces, 2017, 9(18), 15364–15372 CrossRef .
  93. N. Karikalan, et al., Cobalt molybdenum sulfide decorated with highly conductive sulfur-doped carbon as an electrocatalyst for the enhanced activity of hydrogen evolution reaction, Int. J. Hydrogen Energy, 2019, 44(18), 9164–9173 CrossRef .
  94. J. Hu, et al., FeCo2S4 Nanosheet Arrays Supported on Ni Foam: An Efficient and Durable Bifunctional Electrocatalyst for Overall Water-Splitting, ACS Sustain. Chem. Eng., 2018, 6(9), 11724–11733 CrossRef .
  95. B. Dong, et al., Zinc ion induced three-dimensional Co9S8 nano-neuron network for efficient hydrogen evolution, Renewable Energy, 2020, 157, 415–423 CrossRef .
  96. D. Liu, et al., NiCo2S4 nanowires array as an efficient bifunctional electrocatalyst for full water splitting with superior activity, Nanoscale, 2015, 7(37), 15122–15126 RSC .
  97. N. Jiang, et al., Nickel sulfides for electrocatalytic hydrogen evolution under alkaline conditions: a case study of crystalline NiS, NiS2, and Ni3S2 nanoparticles, Catal. Sci. Technol., 2016, 6(4), 1077–1084 RSC .
  98. L. Yang, et al., Amorphous nickel/cobalt tungsten sulfide electrocatalysts for high-efficiency hydrogen evolution reaction, Appl. Surf. Sci., 2015, 341, 149–156 CrossRef CAS .
  99. R.-L. Zhang, et al., Walnut kernel-like iron-cobalt-nickel sulfide nanosheets directly grown on nickel foam: A binder-free electrocatalyst for high-efficiency oxygen evolution reaction, J. Colloid Interface Sci., 2021, 587, 141–149 CrossRef CAS PubMed .
  100. X. Shang, et al., In situ cathodic activation of V-incorporated NixSy nanowires for enhanced hydrogen evolution, Nanoscale, 2017, 9(34), 12353–12363 RSC .
  101. J. Cheng, et al., Hierarchical Ni3S2@2D Co MOF nanosheets as efficient hetero-electrocatalyst for hydrogen evolution reaction in alkaline solution, Fuel Process. Technol., 2022, 229, 107174 CrossRef CAS .
  102. C.-Z. Yuan, One-Step In Situ Growth of Iron–Nickel Sulfide Nanosheets on FeNi Alloy Foils: High-Performance and Self-Supported Electrodes for Water Oxidation, Small, 2017, 13(18), 1604161 CrossRef PubMed .
  103. G. Ren, et al., Ultrafast fabrication of nickel sulfide film on Ni foam for efficient overall water splitting, Nanoscale, 2018, 10(36), 17347–17353 RSC .
  104. J. Jian, et al., Sn–Ni3S2 Ultrathin Nanosheets as Efficient Bifunctional Water-Splitting Catalysts with a Large Current Density and Low Overpotential, ACS Appl. Mater. Interfaces, 2018, 10(47), 40568–40576 CrossRef CAS PubMed .
  105. Z. Chen, et al., Controllable design of nanoworm-like nickel sulfides for efficient electrochemical water splitting in alkaline media, Mater. Today Energy, 2020, 18, 100573 CrossRef CAS .
  106. Q. Ma, et al., Identifying the electrocatalytic sites of nickel disulfide in alkaline hydrogen evolution reaction, Nano Energy, 2017, 41, 148–153 CrossRef CAS .
  107. G. Rahman, S. Y. Chae and O.-s. Joo, Efficient hydrogen evolution performance of phase-pure NiS electrocatalysts grown on fluorine-doped tin oxide-coated glass by facile chemical bath deposition, Int. J. Hydrogen Energy, 2018, 43(29), 13022–13031 CrossRef CAS .
  108. N. Li, et al., Spinel-type oxygen-incorporated Ni3+ self-doped Ni3S4 ultrathin nanosheets for highly efficient and stable oxygen evolution electrocatalysis, J. Colloid Interface Sci., 2020, 564, 418–427 CrossRef CAS PubMed .
  109. S. Zhu, et al., Constructing Stable MoOx–NiSx Film via Electrodeposition and Hydrothermal Method for Water Splitting, Catalysts, 2023, 13(11), 1426 CrossRef CAS .
  110. W.-J. Jiang, et al., Synergistic Modulation of Non-Precious-Metal Electrocatalysts for Advanced Water Splitting, Acc. Chem. Res., 2020, 53(6), 1111–1123 CrossRef CAS PubMed .
  111. C. Liu, et al., P-Doped Iron–Nickel Sulfide Nanosheet Arrays for Highly Efficient Overall Water Splitting, ACS Appl. Mater. Interfaces, 2019, 11(31), 27667–27676 CrossRef CAS PubMed .
  112. T. An, et al., Interlaced NiS2–MoS2 nanoflake-nanowires as efficient hydrogen evolution electrocatalysts in basic solutions, J. Mater. Chem. A, 2016, 4(35), 13439–13443 RSC .
  113. S. Xue, et al., A highly active and durable electrocatalyst for large current density hydrogen evolution reaction, Sci. Bull., 2020, 65(2), 123–130 CrossRef CAS PubMed .
  114. B. Dong, et al., Two-step synthesis of binary Ni–Fe sulfides supported on nickel foam as highly efficient electrocatalysts for the oxygen evolution reaction, J. Mater. Chem. A, 2016, 4(35), 13499–13508 RSC .
  115. X. Wang, et al., Amorphous Multi-elements Electrocatalysts with Tunable Bifunctionality toward Overall Water Splitting, ACS Catal., 2018, 8(11), 9926–9935 CrossRef CAS .
  116. K.-L. Yan, et al., Organic-inorganic hybrids-directed ternary NiFeMoS anemone-like nanorods with scaly surface supported on nickel foam for efficient overall water splitting, Chem. Eng. J., 2018, 334, 922–931 CrossRef .
  117. C. Huang, et al., N-doped Ni-Mo based sulfides for high-efficiency and stable hydrogen evolution reaction, Appl. Catal., B, 2020, 276, 119137 CrossRef .
  118. L. Zhao, et al., Hierarchical Ni3S2–CoMoSx on the nickel foam as an advanced electrocatalyst for overall water splitting, Electrochim. Acta, 2021, 387, 138538 CrossRef .
  119. S. Shit, et al., Cobalt Sulfide/Nickel Sulfide Heterostructure Directly Grown on Nickel Foam: An Efficient and Durable Electrocatalyst for Overall Water Splitting Application, ACS Appl. Mater. Interfaces, 2018, 10(33), 27712–27722 CrossRef PubMed .
  120. S. Shit, et al., Binder-Free Growth of Nickel-Doped Iron Sulfide on Nickel Foam via Electrochemical Deposition for Electrocatalytic Water Splitting, ACS Sustain. Chem. Eng., 2019, 7(21), 18015–18026 CrossRef .
  121. J. Li, et al., Co9S8–Ni3S2 heterointerfaced nanotubes on Ni foam as highly efficient and flexible bifunctional electrodes for water splitting, Electrochim. Acta, 2019, 299, 152–162 CrossRef .
  122. Y. Xiong, et al., Interface-engineered atomically thin Ni3S2/MnO2 heterogeneous nanoarrays for efficient overall water splitting in alkaline media, Appl. Catal., B, 2019, 254, 329–338 CrossRef CAS .
  123. Q. Che, et al., One-step controllable synthesis of amorphous (Ni-Fe)Sx/NiFe(OH)y hollow microtube/sphere films as superior bifunctional electrocatalysts for quasi-industrial water splitting at large-current-density, Appl. Catal., B, 2019, 246, 337–348 CrossRef CAS .
  124. Y. Wu, et al., Novel electrocatalyst of nickel sulfide boron coating for hydrogen evolution reaction in alkaline solution, Appl. Surf. Sci., 2019, 480, 689–696 CrossRef .
  125. X. Shi, et al., Rapid flame doping of Co to WS2 for efficient hydrogen evolution, Energy Environ. Sci., 2018, 11(8), 2270–2277 RSC .
  126. W. Kong, C. Li and W. Wu, A supercritical growth strategy for 1T/2H mixed-phase MoS2 nanosheets of high activity and stability, Int. J. Hydrogen Energy, 2023, 48(81), 31582–31589 CrossRef .
  127. Y. Ito, High Catalytic Activity of Nitrogen and Sulfur Co-Doped Nanoporous Graphene in the Hydrogen Evolution Reaction, Angew. Chem., Int. Ed., 2015, 54(7), 2131–2136 CrossRef PubMed .
  128. M. A. Tekalgne, et al., Hierarchical molybdenum disulfide on carbon nanotube–reduced graphene oxide composite paper as efficient catalysts for hydrogen evolution reaction, J. Alloys Compd., 2020, 823, 153897 CrossRef .
  129. Z. Wang, et al., Co(II)1xCo(0)x/3Mn(III)2x/3S Nanoparticles Supported on B/N-Codoped Mesoporous Nanocarbon as a Bifunctional Electrocatalyst of Oxygen Reduction/Evolution for High-Performance Zinc-Air Batteries, ACS Appl. Mater. Interfaces, 2016, 8(21), 13348–13359 CrossRef PubMed .

This journal is © The Royal Society of Chemistry 2024
Click here to see how this site uses Cookies. View our privacy policy here.