Recent advances in transition-metal-sulfide-based bifunctional electrocatalysts for overall water splitting

Min Wang a, Li Zhang b, Yijia He c and Hongwei Zhu *a
aState Key Lab of New Ceramics and Fine Processing, School of Materials Science and Engineering, Tsinghua University, Beijing 100084, China. E-mail: hongweizhu@tsinghua.edu.cn
bKey Laboratory of Photochemical Conversion and Optoelectronic Materials, Technical Institute of Physics and Chemistry, Chinese Academy of Sciences, Beijing 100190, China
cChina Ship Information Center, Beijing 100192, China

Received 15th December 2020 , Accepted 20th January 2021

First published on 21st January 2021


Abstract

Hydrogen produced via water electrolysis can act as an ideal clean chemical fuel with superb gravimetric energy density and high energy conversion efficiency, solving the problems of conventional fossil fuel exhaustion and environmental contamination. Transition metal sulfides (TMS) have been extensively explored as effective, widely available alternatives to precious metals in overall water splitting. Herein, recent advances, covering preparation methods, intrinsic electrocatalytic performance, and optimization strategies, relating to TMS-based bifunctional electrocatalysts have been summarized systematically and comprehensively. Firstly, a general introduction to the reaction mechanisms and key parameters of the hydrogen evolution reaction (HER) and oxygen evolution reaction (OER) is provided. Next, the physicochemical properties of TMS and typical synthesis methods are introduced to give guidance for fabricating TMS materials with well-defined structures, controllable compositions, and excellent performance. Importantly, the intrinsic activities of TMS-based electrocatalysts and several strategies for improving their bifunctional electrocatalytic performance during water electrolysis are discussed in detail. Finally, perspectives covering the challenges and opportunities related to the further development of TMS-based materials with high activity and long-term durability for overall water splitting are given. The aim herein is to provide guidelines for the design and fabrication of TMS-based bifunctional electrocatalysts with excellent performance and to accelerate their large-scale practical application in water electrolysis.


image file: d0ta12152e-p1.tif

Min Wang

Min Wang is a PhD candidate at the School of Materials Science and Engineering, Tsinghua University, China. She received her B.S. degree in Materials Science and Engineering from Jilin University in 2016. Her research focuses on advanced nanomaterials related to electrocatalysis for energy conversion.

image file: d0ta12152e-p2.tif

Hongwei Zhu

Hongwei Zhu is a Professor at the School of Materials Science and Engineering, Tsinghua University, China. He received his B.S. degree in Mechanical Engineering (1998) and his PhD degree in Materials Processing Engineering (2003) from Tsinghua University. After postdoctoral studies in Japan and the USA, he began his independent career as a faculty member at Tsinghua University (2008 to present). His current research interests involve low-dimensional materials and materials informatics.


1. Introduction

1.1 Background

The growing consumption of conventional fossil fuels and accompanying environmental contamination are affecting global society on an unprecedented scale.1 Consequently, there is a quite imminent need to fundamentally adjust the world energy landscape and build clean, low-carbon, safe, and efficient modern energy systems. As an ideal clean chemical fuel with superb gravimetric energy density and energy conversion efficiency, hydrogen energy is expected to be an excellent candidate to replace traditional fossil fuels.2 Electrochemical water splitting is considered to be one of the most promising hydrogen production technologies, and it can utilize electricity generated via renewable energy sources, such as solar energy, wind power, geothermal energy, and bioenergy, to form a closed loop of renewable energy.3 Unfortunately, the large-scale commercial application of electrochemical water splitting is subject to the following three restrictions: (i) a larger overpotential than the theoretical value (1.23 V) being needed to drive overall water splitting; (ii) the poor stability of electrode materials; and (iii) high cost caused by the scarcity of noble-metal electrocatalysts.4,5 To date, the benchmark electrocatalysts for the hydrogen evolution reaction (HER) and oxygen evolution reaction (OER) are Pt-group- and Ru/Ir-based noble metal materials with small overpotentials and low Tafel slopes.6 These noble metals can suffer from dissolution, agglomeration, and poor durability during the water splitting process.7

The development of low-cost and widely available alternatives with more excellent activities and stabilities than precious metals is urgently needed for overall water splitting, and this is a crucial step in the development of a hydrogen-based economy. There are three main kinds of electrolyzer used for water electrolysis, categorized by the electrolyte: alkaline electrolyzers (AEs), acidic proton exchange membrane (PEM) electrolyzers, and solid oxide electrolyzers (SOEs).8 SOEs are still in the laboratory stage, and they are operated at high temperatures (700–800 °C) to reduce the overpotential. Alkaline electrolyzers have the advantages of simple construction, convenient operation, and low cost. In comparison, water electrolysis carried out in acidic electrolyte involves fewer adverse reactions and higher ionic conductivity, but it demands expensive perfluorinated Nafion-based PEMs.9 As we know, neutral and alkaline electrolytes are favorable for the OER, while acidic electrolytes are beneficial for the HER. Inevitably, this non-negligible mismatch between optimal working environments for the HER and OER has a negative impact on the overall efficiency of water electrolysis systems.10 In this regard, designing and constructing effective bifunctional electrocatalysts that can be employed for both the HER and OER in the same electrolyte is another vital issue related to commercial electrochemical water splitting.

A variety of Earth-abundant and non-precious transition metal (e.g., Mo, W, Ni, Co, and Fe) based compounds have been explored as bifunctional electrocatalysts for overall water splitting, such as transition metal sulfides, phosphides, nitrides, carbides, oxides, and hydroxides.3,10–14 Specifically, transition-metal sulfides (TMS) with distinctive structural features, rich active sites, and adjustable electronic properties and components have attracted widespread research attention,2e.g., layered MoS2 and WS2 and non-layered Ni3S2 and Co3S4. There are multitudinous excellent reviews on research achievements relating to TMS-based electrocatalysts, however, most of them have an emphasis on either one or a limited number of TMS systems for electrochemical water splitting, such as transition-metal dichalcogenides for the HER,15,16 Fe-based electrocatalysts for the OER,17 or Co-based electrocatalysts for overall water splitting.18 Moreover, the use of nanoarchitectonics in TMS-based electrocatalysts to improve their performance for water splitting19 and recent advances relating to metal sulfides, from their controlled fabrication to their applications in electrocatalytic, photocatalytic, and photoelectrochemical water splitting, have also been reviewed.2 However, a targeted review focusing on TMS-based bifunctional electrocatalysts addressing their controllable preparation, intrinsic electrocatalytic activities, and optimization strategies for enhancing their electrocatalytic performance in overall water splitting has not been fully provided to date.

Herein, we focus on recent developments relating to TMS-based bifunctional electrocatalysts, with respect to synthetic methods, intrinsic electrocatalytic activities, and corresponding optimization strategies, aiming to provide a comprehensive overview of the basic information and breakthroughs relating to the use of TMS for electrochemical water splitting. Firstly, the reaction mechanisms for the HER and OER processes together with some significant parameters for evaluating the properties of bifunctional electrocatalysts are introduced. Secondly, the physicochemical properties of TMS and the role of S atoms in electrochemical water splitting are emphatically discussed. Then, typical synthetic methods for TMS are discussed, which exhibit great influence on the electrocatalytic performance. Importantly, the intrinsic activities of TMS-based bifunctional electrocatalysts and several strategies aimed at the optimization of the electrocatalytic performance in overall water splitting are summarized systematically. Possible suitable elements and strategies for improving the electrocatalytic performance of TMS-based bifunctional electrocatalysts are displayed in Fig. 1. Meanwhile, the opportunities and challenges relating to the development of TMS bifunctional electrocatalysts are discussed and an outlook shedding light on directions for further research into TMS is also put forward.


image file: d0ta12152e-f1.tif
Fig. 1 Possible suitable elements and strategies for improving the electrocatalytic performances of TMS-based bifunctional electrocatalysts; transition metals with deeper colors in the periodic table have been the focus of more studies.

1.2 Fundamentals of overall water splitting

1.2.1 Reaction mechanisms. Electrochemical water splitting involves two half reactions, namely the HER and OER. The electrolyzer typically consists of three parts: a cathode, an anode, and the aqueous electrolyte, and a schematic diagram is shown in Fig. 2a. When a certain voltage is applied to the two electrodes, the water reduction reaction occurs at the cathode to produce H2 and the water oxidation reaction happens at the anode to produce O2. The total reaction can be written as follows:
 
2H2O → 2H2 + O2(1)

image file: d0ta12152e-f2.tif
Fig. 2 (a) A schematic illustration of electrochemical water splitting. (b) The general reaction pathways for the HER in alkaline (red lines) and acidic (blue lines) electrolytes. (c) The general reaction pathways for the OER in alkaline (red lines) and acidic (blue lines) electrolytes; the black and purple lines stand for two possible intermediates: M–OOH and M–O, respectively.

The theoretical thermodynamic potential of overall water splitting is 1.23 V under ideal conditions (25 °C and 1 atm). However, a higher operational potential (E) is always necessary to actually carry out the process. The excess potential is named the overpotential (η), which is mainly used to overcome intrinsic activation hindrance at both the anode (ηa) and cathode (ηc), as well as other inevitable obstacles (ηother) due to the electrolyte and contact resistance. Overall:

 
E = 1.23 V + ηa + ηc + ηother(2)

The HER is a two-electron transfer process that happens at the cathode; it is always more active in acidic electrolytes than in alkaline electrolytes because the reduction of protons to H2 is energetically more favorable.20 There are two mechanisms involved the HER, namely the Volmer–Heyrovsky and Volmer–Tafel pathways (Fig. 2b), depending on the way in which adsorbed hydrogen is desorbed from the electrocatalyst.21 In acidic electrolytes, hydronium ions adsorb on the cathode to form hydrogen intermediates (H*) via the Volmer reaction (H3O+ + e → H* + H2O) and then transform into H2via the Heyrovsky reaction (H* + H3O+ + e → H2 + H2O) at low H* coverage or via the Tafel reaction (H* + H* → H2) at high H* coverage. Under alkaline conditions, the HER is more sluggish due to the water dissociation step prior to the formation of H*.22 A water molecule undergoes a dissociation reaction to generate absorbed H* via the Volmer reaction (H2O + e → H* + OH). The desorption process is carried out via the Heyrovsky or Tafel reaction, similar to under acidic conditions.

In the HER process, the free energy of hydrogen adsorption (ΔGH) calculated via density functional theory (DFT) is widely accepted as a descriptor for the interactions between absorbed H* and the electrocatalyst.23 A large negative ΔGH value indicates that the adsorption of hydrogen on the electrode is much easier than desorption, so the Heyrovsky or Tafel reaction is the rate-limiting step. Otherwise, a large positive ΔGH value represents a weak interaction, and the rate-limiting step is the Volmer reaction. Thus, an excellent HER electrocatalyst should possess near-zero ΔGH, with a balance between the absorption and desorption of hydrogen.

The OER is more sluggish than the HER, and it naturally requires a larger overpotential because it involves a complex four-proton transfer process.19 Possible mechanisms for the OER at the anode have been proposed, including almost the same intermediates, such as MO and MOH (M: transitional metal, such as Fe, Co, and Ni).24 The general reaction pathways and widely accepted mechanisms for the OER in alkaline and acidic electrolytes are shown in Fig. 2c. There are two possible pathways for MO intermediates to convert to O2: combining H2O under acidic conditions or OH under alkaline conditions with an MO intermediate to form an MOOH intermediate, which then transforms into O2; and directly uniting two MO intermediates to generate O2. Therefore, the M–O bonds of MOH, MO, and MOOH intermediates play key roles in determining the electrocatalytic OER activity.25 Since the reactions in the OER are all thermodynamically uphill, the reaction with the highest energy barrier will be the rate-determining step.19 The OER rate-determining step can also be determined based on the Tafel slope, and the actual reaction mechanism can be understood further. For example, the Tafel slope is about 120 mV dec−1 when the first reaction (from M to MOH) is the rate-determining step, while the Tafel slope is about 40 mV dec−1 if the second reaction (from MOH to MO) acts as the rate-determining step.

The thermodynamics in each OER step can be studied via DFT calculations. Rossmeisl et al. studied four reactions during the OER and obtained the corresponding Gibbs free energies (ΔGi = ΔGproduct − ΔGreactanteUkBT[thin space (1/6-em)]ln[H+], i = 1–4) under an applied potential (U), where kB is the Boltzmann constant and T is the temperature.26 The largest Gibbs free energy of these is defined as ΔGOER for the OER.3 Under standard conditions (USHE = 0), the relationship between the theoretical overpotential (ηOER) and ΔGOER is as follows:

 
ηOER = (ΔGOER/e) − 1.23 V(3)

The ideal value of ΔGOER is 1.23 eV when ηOER = 0, indicating that the Gibbs free energy in each step is 1.23 eV. However, there is always a gap between ideality and reality due to unsuitable electrocatalyst oxygen bonding forces that can make the OER more sluggish.

1.2.2 Parameters for electrochemical water splitting. There are some important electrochemical water splitting parameters that can be used to evaluate the electrocatalytic performances of materials, such as the overpotential, Tafel slope, exchange current density, turnover frequency, stability, faradaic efficiency, and electrochemically active surface area.

Overpotential (η) is the excess part of the operational potential beyond the equilibrium potential that arises due to an unavoidable intrinsic kinetic barrier.27 As mentioned above, the equilibrium potentials for electrocatalyzing the HER and OER are 0 V and 1.23 V (vs. RHE: reversible hydrogen electrode), respectively. However, applied potentials much higher than the equilibrium potentials are needed to overcome the electrode kinetic barriers in the reaction. Normally, the total activity of an electrocatalyst can be assessed based on polarization curves (after IR compensation) using the geometric current density and overpotential, which are obtained from cyclic voltammetry (CV) or linear sweep voltammetry (LSV) measurements. It is worth noting that different current densities can refer to different overpotential values. Thus, the overpotential at a certain current density (like 10 or 100 mA cm−2) obtained from a polarization curve is one of the most popular descriptors of the total activity.28 The smaller the overpotential at a certain current density, the better the electrochemical activity of the electrocatalyst.

Tafel slope is a key kinetic parameter in electrochemical water splitting, implying how fast the current density increases with increasing overpotential.29 A Tafel plot can be obtained via transforming the current density from a polarization curve into a logarithmic value with base 10 on the x axis, with the corresponding overpotential on the y axis. In a Tafel plot, there is a linear region that can be well-fitted using the Tafel equation:

 
η = a + b[thin space (1/6-em)]log[thin space (1/6-em)]j(4)
where b stands for the Tafel slope, η is the overpotential, j represents the current density, and a is a constant. A smaller Tafel slope always signifies faster charge-transfer kinetics; in other words, increasing the current density leads to a smaller overpotential increase than in a case with a larger Tafel slope. Hence, the Tafel slope can be employed to deduce the possible reaction mechanism, and especially to explain the rate-determining step. When the value of η is equal to zero, the corresponding current density (j) is the exchange current density (j0) that can reflect the intrinsic bonding/charge transfer interactions between the electrocatalyst and reactant. A high j0 value is always a good indication of superior inherent electrocatalyst activity under equilibrium conditions. In short, materials with a small η and a larger j0 value will be efficacious electrocatalysts.

Turnover frequency (TOF) denotes the number of molecules transformed at each active site per unit time, signifying the intrinsic activity of each active site. The TOF can be calculated via the following equation:

 
TOF = (jA)/(αFn)(5)
where j represents the current density at a certain overpotential obtained from the LSV curve; A is the surface area of the electrode; α denotes the electron number of the electrocatalyst (electrons mol−1); F is the Faraday constant (96[thin space (1/6-em)]485.3 C mol−1); and n stands for the molar amount of electrocatalyst. The precise TOF value is difficult to obtain, as it is not possible to accurately determine the number of all accessible active sites involved in the actual reaction. As such, a reasonable approach for calculating the TOF is based merely on either the number of atoms on the surface or the number of easily accessible electrocatalytic sites of the material. Sometimes, the TOF is calculated based on the overall catalytic species existing in the material.11 Despite the imprecision of TOF values, they still play a meaningful role in evaluating the intrinsic catalytic activities of electrocatalysts with similar elements and structures.

Stability is a considerably important parameter for electrocatalysts for practical applications, reflecting the ability of an electrocatalyst to maintain its activity under long-term testing (generally several or tens of hours). In general, there are two electrochemical techniques used to measure the electrocatalytic stability. One is performing CV measurements for thousands of cycles, followed by LSV measurements; the stability can be examined via comparing the LSV curves before and after the cycling CV testing. The other way is to conduct chronoamperometry or chronopotentiometry measurements to show how the potential or current density varies with time; a high retention rate corresponds to good stability. Usually, a current density of 10 mA cm−2 is applied as a standard for evaluating the electrocatalytic stability. To illustrate the potential of an electrocatalyst for use in practical applications, stability at a large current density (e.g., 500 mA cm−2) is absolutely needed.

Faradaic efficiency is defined as the efficiency of electron transfer when catalyzing a desired reaction in the electrochemical system, which is quantified as the experimental to theoretical molar ratio of the numbers of gas molecules. The experimental produced gas can be measured via the traditional water gas displacement method or gas chromatography, while the theoretical value can be calculated via chronoamperometry or chronopotentiometry measurements. A higher faradaic efficiency means less energy loss during the electrochemical reaction.

Electrochemically active surface area (ECSA) is universally utilized to estimate the effective active surface of the working electrode, which is based on the electrochemical double-layer capacitance of the electrocatalytic surface.30 The electrochemical double-layer capacitance can be measured based on the capacitive current associated with double-layer charging in a non-faradaic region from the scan-rate dependence of CV curves.31 Typically, the non-faradaic region is a potential window of 0.1 V centered at the open-circuit potential of the system. A plot of the charging current as a function of the scan rate yields a straight line with a slope equal to the electrochemical double-layer capacitance (CDL). The ECSA of an electrocatalyst sample is calculated via the following equation:

 
image file: d0ta12152e-t1.tif(6)
where Cs is the specific capacitance of the electrocatalyst or the capacitance per unit area of an atomically smooth planar surface of the material, with typical values between 0.015 and 0.110 mF cm−2 in H2SO4 and 0.022 and 0.130 mF cm−2 in NaOH or KOH solution.30 In addition, the electrochemical capacitance can also be acquired from the frequency-dependent impedance based on electrochemical impedance spectroscopy (EIS) studies in the same non-faradaic region.

2. Physicochemical properties of TMS

2.1 Structures and basic properties

Based on crystal structures, TMS can be mainly divided into two categories: layered MS2 and non-layered MxSy. The sulfides of transition-metals in groups 4–7 (M = Mo, W, Ta, Nb, etc.) always crystallize with a graphite-like layered structure, whereas some sulfides of transition-metals in groups 8–12 (M = Fe, Co, Ni, etc.) predominantly adopt non-layered structures. Distinctions between the structures and compositions of these TMS will inevitably bring about differences in their electronic and electrochemical properties.

Typically, each layer of layered MS2 has a thickness of 0.6–0.7 nm, consisting of a hexagonal layer of transition-metal atoms sandwiched between two layers of sulfur (S) atoms.32 These layers, with strong intra-layer covalent bonds between atoms, will be stacked vertically together via relatively weak out-of-plane van der Waals forces. Consequently, layered MS2 can be exfoliated readily into two-dimensional nanosheets.33 Depending on the transition-metal coordination of S atoms and the stacking sequence adopted by multiple layers, layered MS2 can be classified into different phases: 1T, 2H, and 3R, where the digit denotes the number of stacked layers in the crystallographic unit cell and the letter stands for tetragonal, hexagonal, and trigonal lattices, respectively, as shown in Fig. 3a.34 In layered MS2, the transition-metal atoms centered in an octahedral configuration or trigonal prismatic configuration will provide four electrons to fill the bonding states of TMS, so the transition-metal atom is in a +4 oxidation state and the S atom is in a −2 state.


image file: d0ta12152e-f3.tif
Fig. 3 (a) The different metal coordination modes and stacking sequences of layered TMS structural unit cells. The metal can have either octahedral coordination or trigonal prismatic coordination. Octahedral coordination allows stacking sequences with tetragonal symmetry, like AbC, AbC, etc.; trigonal prismatic layers can be stacked in two different ways to form hexagonal symmetry (2H) or rhombohedral symmetry (3R).34 (b and c) The structures of non-layered TMS in pyrite and marcasite phases, in which transition metal atoms and S atoms are shown in orange and yellow, respectively. (d) A side-view of the stable nonpolar pyrite (100) facet as an example of a low-index surface with under-coordinated transition metal cations.38

The electronic properties of TMS are extremely dependent on the filling of the d orbitals of the transition metal. For example, the semiconductor behavior of natural 2H-MoS2 can mainly be ascribed to the fully filled d orbitals of the Mo atom, so 2H-MoS2 is expected to be employed in electronic devices. However, 1T-MoS2, with partially filled Mo atom d orbitals exhibits metallic properties, which will be more suitable for electrocatalysis.35 Interestingly, an MS2 phase transition from 2H to 1T, such as in the cases of MoS2 and WS2, could be realized via the intercalation of alkali metals, which introduce extra electrons and rearrange the d orbitals of metals.36,37

Furthermore, the electrochemical properties of layered TMS also strongly depend on their structures and compositions. The unique stacked structures of layered TMS bring about two distinctive orientations: the basal plane and edge plane, which exhibit anisotropic properties in many aspects. Firstly, the in-plane electrical conductivity of layered MS2 is about 2200 times higher than the interlayer conductivity.34 As a consequence, heterogeneous electron transfer along the edge plane of layered TMS is significantly faster than along its basal plane. As such, the electrochemical activity in the edge plane and basal plane of layered TMS is also predicted to be anisotropic. Typically, sites on the basal planes of group-6 TMS are always inert, while the highly active edge sites mainly contribute to the excellent electrocatalytic performance.39 Taking well-studied MoS2 as an example, 2H-MoS2 exhibits hydrogen adsorption free energy of +0.08 eV on its edge sites compared with 2.00 eV on its inert basal plane.40 Intensive efforts have been made to induce basal plane electrochemical activity. Zheng et al. activated and optimized the basal plane of monolayer 2H-MoS2 for the HER via introducing S vacancies and elastic strain.41 S vacancies could serve as new active sites in the basal plane, as the gap states around the Fermi level would permit hydrogen to bind with exposed Mo atoms directly. What is more, the electrochemical activity of 2H-MoS2 could be optimized further via straining the plane with S-vacancy sites.

In addition, chemical doping with non-metallic or metallic elements is widely considered to be an effective approach for creating active sites in the inert basal plane of MoS2 and for increasing the intrinsic electrical conductivity; e.g., oxygen (O),42 nitrogen (N), phosphorus (P),43 selenium (Se),44 and zinc (Zn)45 elements can be used. Additionally, these dopants can also greatly optimize the hydrogen adsorption free energy, accelerate charge transfer, and, finally, enhance the electrocatalytic performance of MoS2. It is worth mentioning that group-5 TMS can possess active basal-plane sites intrinsically, e.g., in the cases of TaS2 and NbS2. Yakobson et al. proposed that active sites were concentrated in the basal planes and at the edges of H-TaS2 simultaneously according to experimental and theoretical results, and the active basal-plane sites could contribute to self-optimizing behavior during the HER.46

Non-layered MxSy (M = Fe, Co, Ni, etc.) generally shares a pyrite structure or marcasite structure when x = 1 and y = 2, and the conventional unit cells with high similarity are shown in Fig. 3b and c.38 The pyrite structure belongs to the space group Pa[3 with combining macron], in which the transition-metal atoms are located at face-centered cubic (FCC) sites and bonded octahedrally to adjacent S atoms. Each S atom in the pyrite structure is tetrahedrally coordinated to three transition-metal atoms and one S atom, such that an S2 dimer will also be formed.47 The marcasite structure adopts the orthorhombic Pnnm space group, where the body-centered transition-metal atoms are also octahedrally bonded to neighboring S atoms. It can be clearly found that the octahedra in the marcasite structure are edge-sharing, while in the pyrite structure they are corner-sharing. Indeed, the intergrowth or epitaxial growth of the marcasite structure in/on the pyrite structure of non-layered MxSy can achieved readily due to their structural similarities. Moreover, the electronic structures of pyrite-type non-layered MxSy, which predominantly depend on the d-electron count of the transition metal, are diverse, ranging from insulators (such as NiS2) to semiconductors (such as FeS2) to metals (such as CoS2).48 Interestingly, the values of x and y for each particular MxSy structure can be diverse and a series of sulfides can be formed, e.g., NiS, NiS2, Ni3S2, Ni3S4, Ni7S6, and Ni9S8.49 Typically, the composition of non-layered MxSy can also have a great influence on the electronic properties. In nickel sulfides, NiS2 is an insulator while heazlewoodite Ni3S2 shows intrinsic metallic behavior owing to the continuous conductive network, connected via Ni–Ni bonds, in its crystal structure.50

Differences in the crystal structures and compositions of layered MS2 and non-layered MxSy result in discrepant electronic and electrochemical properties. However, the main strategies for improving the electrocatalytic performance are comprehensive, and they involve enhancing the intrinsic activity of each active site, increasing the number of active sites, and improving the electrical conductivity.1,51 Edge engineering and defect engineering have been developed to expose or create more active sites, and strain engineering, facet engineering, heteroatom doping, and composite designing have been explored to accelerate electron transportation in these electrocatalysts during water splitting. Details of these optimization approaches for electrocatalytic performance will be discussed below.

2.2 The role of S atoms in the HER and OER

Revealing the true active sites or species is crucial for developing new electrocatalysts aimed at efficient overall water splitting. Although the main active sites in TMS are always metal atoms, understanding the role of S atoms in TMS is still an important task. Basically speaking, S atoms in TMS can contribute to special structures and abundant species, such as in the 2D layered structures of MS2 (M: transition metal in group 4–7), and in the nickel sulfides Ni3S2, NiS2, and Ni3S4. The roles of S atoms in catalyzing the HER and OER can be divided into two types: a direct role and an indirect role, depending on the contribution of S atoms to the electrocatalytic performance.

In the HER, some exposed S atoms and S vacancies on the edges or on the basal plane of TMS can have a great effect on the electrocatalytic performance; this can be thought of as an embodiment of the direct role of S atoms. Based on experimental and theoretical results, Jaramillo et al. identified that the active sites of MoS2 during the HER could be S atoms on the edges under low H coverage (1/4 ML).52 What is more, S vacancies on the basal plane of layered 2H-MoS2 could contribute to enhanced HER activity,41 which is a powerful method for improving the HER performance (Fig. 4a). On one hand, S vacancies can change the coordination environment of adjacent metal atoms: metal atoms with low coordination will be more active and will possess higher intrinsic electrocatalytic activity. On the other hand, S vacancies can induce more active sites and significantly increase the number of active sites. Therefore, S vacancies can create new active sites on the inert basal plane of 2H-MoS2, where gap states around the Fermi level can allow hydrogen binding directly with exposed Mo sites. Furthermore, ΔGH can be finely manipulated via changing the number of S vacancies (Fig. 4b).


image file: d0ta12152e-f4.tif
Fig. 4 (a) Schematic illustrations showing top and side views of MoS2 with S vacancies on the basal plane. (b) A plot of calculated free energy versus the reaction coordinate of MoS2 during the HER over an S vacancy range of 0–25%.41 (c) A schematic diagram of the HER reaction mechanism of TMS when metal atoms act as the active sites. (d) The obtained structures of H2O, activated H2O, OH, and H intermediates during water dissociation and H adsorbed on the Ni3S2 (003) surface after relaxation.50

The indirect role of S atoms in the HER mainly arises through providing a place for hydrogen adhesion and separation when metal atoms act as the active sites (Fig. 4c). When the HER took place on Ni3S2 nanoporous thin films with exposed (003) facets, Dong et al. came to the following conclusions based on theoretical calculations: H2O was preferentially adsorbed on Ni sites via the Volmer process, and S sites were more suitable for the formation of H2 in the Tafel stage (Fig. 4d).50 More straightforward evidences for the indirect role of S atoms has been found in experiments by Staszak-Jirkovsky et al.4 They proposed that TMn+ (Co2+/Mo4+) plays a key role in accelerating the sluggish water dissociation process in alkaline solutions, while the recombination of absorbed hydrogen could happen at S sites. Due to the amorphous phases of the catalysts, they further mimicked Sδ–TMn+ building blocks in chalcogels via a simple model based on Sδ–Cn+–H2O clusters to verify the indirect promotion of S sites, where Cn+ represents K+/Ba2+ cations. In this case, S sites were proved to be enhancers of the HER activity of metal sites in alkaline solutions.

The OER is more complex than the HER, involving the formation of many intermediates during the reaction, such as –OH and –OOH intermediates. Subbaraman et al. proposed that the increasing OER activities of 3d-M hydr(oxy)oxides (M = Fe, Co, Ni, or Mn) are closely connected with the decreased bond strength of OHad–Mn+, which was mainly ascribed to enhanced repulsion between the metal d-band and coordinated oxygen p-band centers.53 As we know, electronegative S atoms in TMS would hinder the coordination of –OH with metal atoms due to repulsion between the S 3p orbital and O 2p orbital. However, this 3p–2p repulsion could accelerate the further oxidation of –OOH intermediates, and the formation of –OOH intermediates could also be promoted by the delocalized electrons of S atoms.3

There also have been numerous efforts to reveal the true role of the S atoms of TMS in the OER. As a general rule, TMS tend to function as pre-catalysts in the OER because of the conversion of sulfides to transition metal hydr(oxy)oxides during the OER process.54 Zhang et al. confirmed that Ni-Fe (oxy)hydroxides that formed in the near-surface region of Fe-doped Ni3S2 on Ni foam (Fe-Ni3S2/NF) are the real active materials that actually contribute to the OER activity.8 In this case, the bifunctional Fe-Ni3S2/NF electrode could act both as a catalyst for the HER and also as a pre-catalyst for the in situ formation of active-oxidation-state species for electrocatalyzing the OER, which was verified based on high-resolution XPS spectra before and after long-term testing (Fig. 5). No obvious changes appeared in the component types and binding energies of both Ni and S elements, indicating that Fe-Ni3S2/NF possesses superior durability in terms of structure and composition during the HER process. Nevertheless, high-resolution S 2p XPS spectra revealed that the signal intensity of S on the surface was obviously reduced after OER testing and even after further Ar+ etching, suggesting that near-surface sulfides had been transformed to Ni-Fe (oxy)hydroxides in the OER process.


image file: d0ta12152e-f5.tif
Fig. 5 (a) A schematic diagram of Fe-Ni3S2/NF applied to overall water splitting in a two-electrode configuration. (b) The polarization curves of an Fe17.5%-Ni3S2/NF couple and Pt/C/NF//IrOx/NF water electrolyzers in alkaline solution at 5 mV s−1. (c) Ni 2p and (d) S 2p high-resolution XPS spectra of Fe17.5%-Ni3S2/NF before and after long-term HER testing. (e) S 2p high-resolution XPS spectra of Fe17.5%-Ni3S2/NF before and after long-term OER testing. Ar+ etching was employed to remove surface materials.8

Coincidentally, Mabayoje et al. discovered that NiS can transform into amorphous NiOx during the OER, which was derived from the finding that S anions in NiS were completely depleted in the active form of the electrocatalyst.55 In the Ni 3p XPS region, they speculated that the number of oxidation states of Ni in amorphous NiOx induced by Ni vacancies are higher than in crystalline NiO. Furthermore, amorphous NiOx derived from NiS had a relatively high electrochemically active surface area and exhibited enhanced OER activity. Thus, TMS can serve as pre-catalysts to produce more active transition metal hydr(oxy)oxides for electrocatalyzing the OER.

In conclusion, when TMS are employed as electrocatalysts, the main active sites are always metal atoms, while S atoms play pivotal roles in determining the catalytic activity. In the OER process, S atoms can promote the formation of more active transition metal hydr(oxy)oxides. In the HER process, exposed S atoms can act as active sites under specific conditions, providing places for hydrogen adhesion and separation when metal atoms are the active sites, while S vacancies can induce more active sites and change the coordination environments of adjacent metal atoms, enhancing the activity. A clear understanding of the roles of the S atoms/vacancies of TMS in water splitting could bring enlightenment for designing novel electrocatalysts. Nevertheless, more systematic research revealing the exact roles of these atoms in actual reactions should be conducted, and it could greatly accelerate the development of TMS for water electrolysis.

3. Typical synthetic methods for TMS

The controllable synthesis of TMS with well-defined shapes, sizes, hierarchical structures, and defect states is the initial step in obtaining high-performance bifunctional electrocatalysts. Up to now, various methods have been developed for preparing high-quality TMS, which can be divided into two types, ‘bottom-up’ and ‘top-down’ strategies.2 A schematic illustration of typical synthetic methods for TMS materials is shown in Fig. 6.
image file: d0ta12152e-f6.tif
Fig. 6 A schematic illustration of typical synthetic methods for TMS materials.

3.1 ‘Top-down’ strategies

Typically, top-down strategies refer to the exfoliation or successive reduction of a bulk starting material to target nanostructures, such as via ball milling and sputtering methods. In particular, exfoliation is a simple, low-cost, and promising technique for fabricating high-quality, layered low-dimensional TMS materials; mechanical exfoliation and liquid phase exfoliation are common methods.56

Mechanical exfoliation is a clean and easy way to gain pristine TMS of high-quality that are suitable for fundamental studies and potential applications based on the intrinsic thickness-dependent properties of TMS. Typically, bulk TMS materials with layered structures are applied as starting materials, and TMS parts are peeled off using adhesive tape (‘sticky’ tape) and then transferred onto the target surface.2 The TMS parts attach to the surface via van der Waals forces. Li et al. produced single-layer and multilayer MoS2 nanosheets via mechanical exfoliation.57 However, some challenges still remain in relation to mechanical exfoliation due to the following points: the bulk starting crystals are expensive; a transfer process is essential for further applications; and the as-exfoliated TMS flakes always have random shapes and low yields.

Liquid phase exfoliation is an efficient method for exfoliating TMS in solution; the yields are considerably higher than those obtained via mechanical exfoliation, while the product is of lower quality. There are two approaches for exfoliating TMS in the liquid phase. The first one is exfoliation via a mechanical method, such as shearing, sonication, stirring, grinding, or bubbling. Exfoliation mechanisms are mainly based on external force induced via mechanical effects and interactions with solvent molecules.58 Coleman et al. reported that layered compounds such as MoS2 and WS2 could be easily exfoliated into individual layers via sonication exfoliation in solution (Fig. 7a and b).59 This procedure was insensitive to water and air and could be potentially scaled up to obtain large quantities of exfoliated products. Additionally, hybrid dispersions or composites could be made for further film-forming via blending these materials with suspensions of other nanomaterials or polymer solutions. Later, Varrla et al. proposed the large-scale shear-exfoliation of MoS2 nanosheets in aqueous surfactant solutions using a kitchen blender (Fig. 7c and d).60Via optimizing the processing parameters, such as the MoS2 concentration, mixing time, liquid volume, and rotor speed, high concentrations (0.4 mg mL−1) and high production rates (1.3 mg min−1) could be achieved. In addition, the surfactant concentration not only influenced the nanosheet concentration but it also had great effects on the nanosheet lateral size and thickness, with higher surfactant concentrations giving smaller flake sizes. In this regard, the average MoS2 nanosheet dimensions could be controlled during shear exfoliation, at least in the ranges of about 40–220 nm for length and about 2–12 layers for thickness.


image file: d0ta12152e-f7.tif
Fig. 7 (a) Photographs of dispersions of MoS2 and WS2 in NMP. (b) TEM images of MoS2 and WS2 flakes produced via sonication exfoliation.59 (c) A photograph of a kitchen blender. (d) TEM images of typical MoS2 flakes produced via shearing exfoliation.60 (e) A schematic diagram of the preparation of trilayer MoS2 nanosheets via chemical intercalation. (f and g) TEM images of the exfoliated MoS2 nanosheets produced via chemical intercalation.62 (h) A schematic illustration of the mechanism of MoS2-flake formation via electrochemical intercalation. (i and j) TEM and HRTEM images of exfoliated MoS2 flakes produced via electrochemical intercalation.63

As reported, the surfactant plays a key role in liquid-phase exfoliation, and it could greatly influence the dimensions of the products. Zhao et al. prepared WS2 nanodots with high quality and uniformity that were dispersed in aqueous solution via liquid-phase exfoliation from bulk crystals with the aid of ultrasonication in surfactant aqueous solution.61 The size distributions of WS2 nanodots showed that the lateral sizes were 2.7 ± 0.8 nm and the height was 0.7 nm, which was about the size of a single layer. Remarkably, a high concentration of 1T-WS2 existed in the as-prepared WS2 nanodots, which could be responsible for its enhanced electrocatalytic activity.

The second liquid-phase exfoliation route involves ionic intercalation, where lithium ions are typically intercalated between the TMS layers to reduce van der Waals forces between layers via enlarging the interlayer spacing.64 Ionic intercalation is beneficial for subsequent exfoliation, and it can involve chemical intercalation and electrochemical intercalation. Fan et al. developed an chemical intercalation method via the incorporation of sub-stoichiometric amounts of n-butyllithium (BuLi) to controllably exfoliate MoS2 crystals into trilayer nanosheets (Fig. 7e–g).62 The intercalated lithium at the edges of the MoS2 crystals could serve as a wedge to promote the exfoliation of MoS2via solvent molecules with a yield of 11–15 wt% exfoliated nanosheets in ethanol/water mixtures. The exfoliation efficiency of a pre-intercalated sample was notably increased by at least 1 order of magnitude compared with the starting MoS2 microcrystals. However, some shortcomings related to chemical intercalation are inevitable, like extreme sensitivity to ambient conditions and the small lateral size of the chemically exfoliated nanosheets. Liu et al. prepared large-area atomically thin MoS2 nanosheets via the electrochemical exfoliation of a bulk MoS2 crystal (Fig. 7h–j).63 Firstly, ·OH, ·O radicals, and/or SO42− anions were inserted into the MoS2 layers under a positive bias applied at the working electrode, weakening the van der Waals interactions between layers. Then, the oxidation of these radicals and/or anions could lead to the release of O2 and/or SO2 gas, which could cause the great expansion of the MoS2 interlayers. Finally, atomically thin MoS2 flakes were detached from the bulk MoS2 crystal by the erupting gas and then suspended in solution.

Ball milling is a low-cost and easy scalable mechano-chemical route for preparing TMS nanostructures. In the ball-milling process, powder particles undergo severe mechanical deformation due to collisions with stainless-steel balls and the vessel; they then become continually deformed, cold-welded, and fractured. Concurrently, mechano-chemical reactions and/or solid-state reactions can happen in powder blends. The ball-milling time is a key parameter, and an inert gas atmosphere is usually used to protect the TMS powder from oxidation. Homogeneous NiS powders were prepared by ball milling of several mixed phases, such as Ni3S2, Ni7S6, NixS6, and Ni3S4, up to 12 h under an Ar atmosphere.65 Similarly, Ambrosi et al. fabricated a MoS2 electrocatalyst via ball milling bulk natural MoS2, and the size reduction contributed to the enhanced electrocatalytic performance.66

3.2 ‘Bottom-up’ strategies

Proverbially, bottom-up strategies refer to the build-up of nanomaterials atom by atom, molecule by molecule, or cluster by cluster, for example via hydrothermal method, solvothermal method, chemical vapor deposition, electrochemical deposition, atomic layer deposition, thermal deposition, and solid-phase synthesis.

A hydrothermal method is a chemical reaction in water at both high temperature and under high pressure, which usually happened in a sealed pressurized vessel. This method is a simple yet efficacious way to fabricate TMS nanomaterials with controlled compositions and structures, and it can be easily realized via adjusting the metal salt precursors and temperature/pressure conditions. Luo et al. fabricated NiS2 hollow microspheres using a hydrothermal approach with Ni(NO3)2 and Na2S2O3 solutions as the Ni and S sources, respectively.67 In this typical procedure, the precursors were dissolved in water and stirred with ethylene glycol and oxalic acid for 20 min before being transferred into a Teflon-lined stainless-steel autoclave, then heated at 180 °C for 12 h in an electric oven and cooled to room temperature naturally. Finally, the product was washed with deionized water and ethanol and dried overnight in a vacuum oven at 60 °C. Interestingly, when annealed under a mixed atmosphere, the obtained NiS2 hollow microspheres could further be transformed into NiS porous hollow microspheres (Fig. 8a–c). Interestingly, Dong et al. synthesized hierarchical CoS with flower-like, ball-like, cube-like, and hollowed-out nanostructures via the one-pot hydrothermal reaction of Co(NO3)2·6H2O and thiourea.68 Thiourea played an important role in the reaction, as it decomposed to produce S2− for the formation of CoS and served as a structure-directing agent to control the growth of CoS crystals. The morphologies of the CoS nanostructures formed could be modified via fine-tuning the molar ratios of the solvents and reactants, the reaction temperature, the reaction time, and the ligand type (Fig. 8d–i). The growth process of the flower-like CoS nanostructures could be divided into three stages: the formation of small-sized nanoflakes; the formation of a skeleton sphere via the attachment of nanoflakes through an Ostwald ripening mechanism; and the development of hierarchical flower-like nanostructures via a dissolution–recrystallization process.


image file: d0ta12152e-f8.tif
Fig. 8 (a) A schematic diagram of the hydrothermal synthesis of NiS2 and NiS porous hollow microspheres. (b) An SEM image of NiS2 hollow microspheres. (c) An SEM image of NiS porous hollow microspheres.67 (d–i) FESEM images of CoS synthesized at 160 °C for given reaction times of 2 h, 5 h, 12 h, 24 h, 36 h, and 48 h, respectively.68

Due to the ease of regulation of the reaction, effective element doping can be achieved easily using the hydrothermal method. Xiong et al. prepared Co-doped MoS2 nanosheets with different Co content levels via a facile one-step hydrothermal approach.69 During the reaction, (NH4)6Mo7O24·4H2O, thiourea, and Co(NO3)2·6H2O were applied as the precursors for the hydrothermal preparation of MoS2 nanosheets covalently doped with Co. The level of Co doping could be regulated simply via changing the added amount of Co(NO3)2·6H2O precursor. Remarkably, a hierarchical flower-like structure with improved conductivity was formed, composed of thin nanosheets, which could be employed as a superior bifunctional electrocatalyst for overall water splitting.

The solvothermal method is very similar to the hydrothermal method, except that an organic solution is applied as the precursor medium instead of water. Li et al. prepared MoS2 nanoparticles on reduced graphene oxide (RGO) sheets via the one-step selective solvothermal reaction of (NH4)2MoS4 and hydrazine in an N,N-dimethylformamide (DMF) solution of graphene oxide (GO) sheets at 200 °C (Fig. 9).70 During this process, the (NH4)2MoS4 precursor was reduced to MoS2 nanoparticles that were laid uniformly and flatly on RGO that was formed from GO upon hydrazine reduction. Importantly, the GO sheets acted as a substrate for the nucleation and subsequent growth of MoS2 nanoparticles, avoiding MoS2 coalescing into 3D particles of various sizes. The selective growth of MoS2 on GO could be attributed to interactions between the Mo precursor and the functional groups of the GO sheets in a suitable solvent environment, because when DMF was replaced with H2O as the solvent, only two separate phases of MoS2 and RGO could be obtained. Yuan et al. reported the direct growth of Fe-doped Ni3S2 nanosheet arrays on a conductive Fe–Ni alloy foil via a one-step solvothermal method.71 A piece of Fe–Ni alloy foil was immersed into a mixture of Na2S solution and ethanol, and this was then transferred into a Teflon-lined stainless-steel autoclave under N2 flow. The reaction was carried out at 200 °C for 12 h in an electric oven and the obtained Fe-doped Ni3S2 nanosheet arrays grown in situ on the surface of the alloy foil could be applied as a self-supported electrode for water oxidation.


image file: d0ta12152e-f9.tif
Fig. 9 (a and b) Schematic diagrams of the solvothermal synthesis of MoS2 in solution with and without graphene sheets. (c and d) SEM and (insets) TEM images of the MoS2/RGO hybrid and free MoS2 particles.70

As key reaction factors in the solvothermal method, the temperature and precursor concentration exhibit great effects on the structural modification of CoS architectures, as proved by Wang et al.72 3D flower-like CoS architectures and CoS microspheres were synthesized through a one-pot solvothermal method using CoCl2·6H2O and thioacetamide as the precursors. According to crystal growth theory, the nucleation and subsequent growth of CoS could be regulated via adjusting the supersaturation of the reaction solution, which is related to the precursor concentration and temperature. CoS microspheres could be formed at a high starting precursor concentration, while flower-like CoS architectures were obtained at a low precursor concentration due to limited homogenous nucleation. Likewise, the higher the reaction temperature, the faster the nucleation rate and crystal growth rate. As a result, the controllable fabrication of various CoS nanostructures could be realized via a solvothermal route via synergistically adjusting the temperature and precursor concentrations, providing a promising way to synthesize TMS with specific morphologies and sizes.

Chemical vapor deposition (CVD) is defined as the deposition of solid materials onto a heated substrate from the vapor phase after several chemical reactions, which is well suited to the preparation of TMS thin films or nanoflakes with high crystallinity on a substrate. During the reaction, solid-state precursors of metal compounds or elemental metal will react with the sulfur source H2S or vaporized S at high temperatures. Yuan et al. reported a facile CVD method to synthesize hexagonal 1T-VS2 single-crystal nanosheets.73 Typically, a vanadium chloride (VCl3) precursor in an Al2O3 crucible was placed in a quartz tube at the center of the heating zone of a furnace. A SiO2/Si substrate was placed on top of the crucible to collect the product. Sulfur was placed upstream of the quartz tube and it migrated with the carrier gas, a mixture of N2 and 10% H2, at ambient pressure. The furnace was gradually heated to 550 °C at a heating rate of 20 °C min−1 and it was cooled to room temperature naturally after being maintained at high temperature for 10 min. Well-aligned hexagonal and nearly “half-hexagonal” VS2 nanosheets mostly parallel to the substrate could be obtained via tuning the addition of precursors.

Furthermore, in CVD, the reaction pressure and substrate both have great influence on the growth of TMS. Samad et al. directly synthesized phase-pure FeS2 thin films with micron-sized crystalline domains via atmospheric pressure CVD with metallic CoS2 film as the substrate.74 The synthesis of FeS2 was achieved via reacting di-tert-butyl disulfide (TBDS) and FeCl3 precursors at 400–450 °C. CoS2 has the same cubic crystal structure as FeS2 with only 2% lattice mismatch between them, so it is highly desirable as a film growth substrate that is unaffected by the sulfur atmosphere and as a conductive substrate for potential electrochemical applications. Subsequently, Shi et al. fabricated thickness-tunable TaS2 flakes and ultrathin centimeter-sized films on gold (Au) foil via low-pressure CVD (LPCVD) and atmospheric-pressure CVD (APCVD) routes with TaCl5 and S as precursors.75 Replacing the general SiO2/Si substrate for TMS growth with Au foil had great advantages, e.g., Au foil could be applied as a metal catalyst for preparing TaS2 and a conductive current collector of TaS2 electrode, providing an opportunity to probe either fundamental physical phenomena or applications related to dimensionality effects. Intriguingly, TaS2 tended to grow into centimeter-sized uniform, several-layer-thick 2H-TaS2 thin films via the LPCVD route, while APCVD was an effective approach to grow TaS2 with tailored thicknesses due to the excess precursor feeding rate in the synthesis procedure. Schematic illustrations of these two growth processes and the as-prepared TaS2 are shown in Fig. 10.


image file: d0ta12152e-f10.tif
Fig. 10 (a) A schematic illustration of the LPCVD growth process of TaS2. (b) An SEM image of 2H-TaS2 flakes (synthesized at 750 °C for ∼10 min) on Au foil. (c) An SEM image of 2H-TaS2 flakes (synthesized at 750 °C for ∼20 min) on Au foil, and an AFM height image of a transferred film edge. (d) A schematic illustration of the APCVD growth process of TaS2. (e and f) SEM images of 2H-TaS2 domains on Au foil with growth times of 3 and 5 min. (g) A plot of the tunable thickness of 2H-TaS2 as a function of the growth time.75

For further optimizing the preparation conditions for the CVD growth of TMS, novel methods originating from CVD have been investigated in recent years, like plasma-enhanced CVD (PECVD) and metal–organic CVD (MOCVD). The reaction temperature could be lowered to 150–300 °C using a PECVD technique, thereby making it possible to deposit active TMS on plastic and flexible substrates. Kim et al. prepared uniform wafer-scale 1T-WS2 film via a PECVD technique.76 Different from common CVD methods, the plasma in an ionized gas state with ions and radicals does not have high thermal energy. The growth temperature was maintained at 150 °C, enabling the direct preparation of uniform and stable 1T-WS2 films on both flexible polymer and rigid dielectric substrates. Specifically, organic–metallic precursors are used in MOCVD, and this method can be applied in modern semiconductor industries to prepare single-crystal epitaxial films. Song et al. prepared transferable few-layer MoS2 structures with desired sizes and patterns via an MOCVD method with a Au catalyst.77 The MOCVD reaction encompassed the following steps: the low-temperature reaction of [Mo(CO)6] and a pre-deposited Au thin film to form a Mo–Au surface alloy; and the reaction of the Mo–Au surface alloy with H2S to obtain large-scale MoS2 atomic layers that could be isolated from the substrate via etching the Au films. Importantly, the MoS2 atomic layers could be fabricated with desired patterns through simply defining the pattern of the Au thin film via conventional lithographic techniques.

Solid-phase synthesis has been widely used to convert metal or metal compounds directly to TMS via a sulfurization process at high temperatures; it is also a promising method for fabricating self-supported TMS electrodes. Wang et al. reported hierarchical WS2 film that was grown in situ on tungsten foil via one-step solid-phase synthesis, namely a surface-assisted chemical vapor transport method (Fig. 11a and b).78 Tungsten foil and sulfur powder were sealed together in a small quartz ampoule under high vacuum and reacted at high temperature for 10 min to get thin WS2 film with a hierarchical structure on the surface of the tungsten foil. Tungsten foil acted as the raw material for solid-phase synthesis and as the substrate for the nucleation and growth of WS2 simultaneously. The strong interaction between WS2 film and tungsten foil dramatically enhanced the electrical conductivity and electrocatalytic durability of this self-supported WS2 electrode. Jin et al. fabricated vertically oriented single-crystal FeS2 nanowires from steel foil via solid-phase synthesis.79 During the sulfurization procedure, low-carbon steel foil, placed near the hot zone of the tube, was heated at 350 °C for 2 h under a sulfur atmosphere that was formed via evaporation from upstream sulfur powder.


image file: d0ta12152e-f11.tif
Fig. 11 (a) Schematic diagrams of WS2 film on tungsten foil prepared via the SACVT method. (b) An SEM image of the WS2 film.78 (c–h) SEM images of FeS2, CoS2, NiS2, PyS2, (Fe0.48Co0.52)S2, and (Co0.59Ni0.41)S2 thin films on graphite disk substrates prepared via solid-phase synthesis.80

What is more, metal sources can be deposited onto other substrates to reduce the utilization of metal. Faber et al. prepared pyrite-phase FeS2, CoS2, and NiS2 thin films and thin films of their alloys via a simple solid-phase synthesis reaction (Fig. 11c–h).80 Firstly, high-purity metal was deposited on the polished surface of a graphite disk substrate via electron-beam evaporation. To prepare ternary pyrite alloy films, metal bilayers containing two kinds of metals were sequentially evaporated onto graphite disk substrates. Finally, the metalized substrates were loaded into the center of a fused silica tube and heated at 500 °C for 1 h under a sulfur atmosphere until they were converted into their corresponding disulfides. In particular, solid-phase synthesis could also be combined with other methods, such as hydrothermal, solvothermal, and sol–gel methods. Miao et al. prepared mesoporous FeS2 materials with large surface areas using a sol–gel method followed by solid-phase synthesis treatment.81 Amorphous Fe2O3 was first obtained via an inverse micelle sol–gel method, and it was then sulfurized to FeS2 using H2S gas and sulfur powder as sulfur sources.

Except for the direct growth of TMS materials on substrates, TMS powder can also be synthesized via solid-phase synthesis. Single-crystal FeS2 nanorods, nanobelts, and nanoplates have been prepared via the solid-phase synthesis reaction of FeCl2 or FeBr2 with sulfur powder at high temperatures.82 During solid-phase synthesis reactions, dehydrated FeCl2 or FeBr2 precursor powder was heated at 425 °C under a flowing sulfur atmosphere to produce phase-pure cubic iron pyrite. Interestingly, Bara et al. investigated the effects of the degree of sulfurization on the preparation of MoS2 slabs on an α-Al2O3 substrate from MoO3 upon reaction with a H2S/H2 gas phase.83 The degree of sulfurization during solid-phase synthesis strongly depended on the surface orientation of the α-Al2O3 substrate, and this had a clear impact on the strength of the metal–support interactions in the oxide states with consequences for the sulfurization, stacking, size, and orientation of MoS2 slabs. Eventually, the maximum degree of sulfurization was about 90% at low temperature (300 °C) for the C (0001) plane with a larger MoS2 slab size. However, the solid-phase synthesis process in this case was similar to the CVD method.

In summary, numerous effective preparation methods have been established for the fabrication of high-quality TMS materials, which could be divided into ‘bottom-up’ and ‘top-down’ strategies. From the viewpoint of surface morphology, the obtained TMS exhibited different morphologies, e.g., nanodots, nanoparticles, nanosheets, flower/ball/cube-like nanostructures, microspheres, and thin films, which strongly depended on the synthesis method and conditions used. The experimental parameters all play key roles in the controllable synthesis of TMS materials, such as the sulfur/metal source type, the precursor concentration, the solvent type, and the reaction temperature/time. Importantly, self-supported TMS electrodes could be fabricated via the sulfurization of metal substrates. Although various methods have been successfully employed to prepare TMS materials, the growth mechanisms of TMS materials with different morphologies and compositions remain unclear; these require more investigation in order to deeply understand the underlying behavior.

4. Water splitting properties of TMS

Numerous efforts have been made to develop efficient and inexpensive bifunctional electrocatalysts for electrochemical water splitting. Of these new bifunctional electrocatalysts, TMS have attracted extensive research attention due to their excellent electrocatalytic performance, wide availability, and low cost. Countless TMS have been explored and their great potential for efficiently catalyzing the HER or OER has been revealed. Nevertheless, most of these TMS electrocatalysts can only be utilized for either the HER or OER because of their intrinsic electrochemical properties; this can restrict their electrocatalytic activity and durability, with strong dependence on the pH of the electrolyte. For instance, most HER electrocatalysts are more active in acidic electrolytes, while most effective electrocatalysts for the OER exhibit high solubility in acidic electrolytes and thus are restricted to alkaline media. Consequently, discovering and designing bifunctional electrocatalysts with high activity and durability for both the HER and OER in the same electrolyte is technologically important. This section comprehensively reviewed the intrinsic electrocatalytic performance and optimizing strategies of layered TMS (M = Mo, W, Ta, etc.) and non-layered TMS (M = Fe, Co, Ni, etc.) for overall water splitting.

4.1 Bifunctional layered TMS (MS2: Mo, W, Ta, etc.) electrocatalysts

4.1.1 Intrinsic electrocatalytic performance. Layered TMS with versatile and tunable properties have been extensively studied as robust substitutes for noble metals for electrocatalyzing the HER. For example, MoS2,84 WS2,78 TaS2,85 and NbS2 (ref. 86) all can be applied as HER electrocatalysts with excellent activity and stability. Lin et al. chemically unzipped WS2 nanotubes to make metallic WS2 nanoribbons with increased edge active site content and high electrical conductivity.87 The enhanced electrocatalytic HER activity of WS2 is strongly associated with its well-exposed edge defects. Moreover, Lukowski et al. exfoliated metallic 1T-MoS2 nanosheets via lithium intercalation from semiconducting 2H-MoS2 nanostructures.36 The metallic 1T-MoS2 nanosheets boosted the electrocatalytic HER activity with a low overpotential of −187 mV at 10 mA cm−2 and a small Tafel slope of 43 mV dec−1, showing facile electrode kinetics, low-loss electrical transport, and a proliferated density of active sites. While the basal plane of 2H-MoS2 is nonreactive, the basal plane of 1T-MoS2 is active; this mainly arises from the affinity of the surface S sites for binding H species.88

Hence, the well-exposed edge sites and good electrical conductivity properties of layered TMS mean they show remarkable intrinsic electrocatalytic HER activities. However, these kinds of TMS are not intrinsically suitable as OER electrocatalysts, and little research has been reported in which they could be directly used for electrocatalyzing the OER. In a rare example, pure 2H-TaS2 nanoflakes were reported to have the potential to electrocatalyze both OER and HER processes with overpotentials of 315 mV and 145 mV, respectively, at 10 mA cm−2,89 and MoS2 microspheres on conductive Ni foam, with enhanced conductivity and accelerated electron transfer, showed great potential for OER applications.90 More effort should be devoted to expanding the application of these layered TMS to the OER.

4.1.2 Strategies for optimizing the electrocatalytic performance of layered TMS. To make these layered TMS more suitable for both the HER and OER, numerous efforts have been put forward to optimize their electrochemical properties via improving the intrinsic electrical conductivity, enhancing the chemisorption capabilities for O-containing species, and/or improving stability. Practically, three main strategies are used to achieve this goal: phase engineering, composite designing, and heteroatom doping.

Phase engineering is a promising way to regulate the electrocatalytic OER activities of layered multiphase TMS. Traditionally, phase transformation can be realized via various strategies, such as chemical exfoliation, ion intercalation, strain engineering, and heteroatom doping. Wu et al. exfoliated 2D MoS2 and TaS2 nanosheets obtaining both 2H and 1T polymorphs and explored their electrocatalytic OER performances in an acidic medium.91 The best OER performance, which was comparable to a benchmark IrO2 electrocatalyst, came from 1T-MoS2, followed by 1T-TaS2, 2H-MoS2, and 2H-TaS2, unraveling the polymorphic dependence of the electrocatalytic OER activity. A theoretical study also revealed the same electrocatalytic activity trend, in which the 1T polymorph showed higher activity than the 2H layered TMS counterpart. Additionally, Wei et al. found that the spontaneous phase transformation of MoS2 from the 2H to 1T phase was triggered by strong metal–support interactions during iridium (Ir) adsorption on the surface of 2H-MoS2.92 Strong charge transfer between MoS2 and the adsorbed Ir atoms could promote the phase transformation of MoS2 and enhance the electrocatalytic activity for overall water splitting, which was superior to even commercial Pt/C and IrO2 electrodes. The significantly improved electrocatalytic performance could be ascribed to enhanced electrical conductivity and hydrophilicity, increased numbers of interfaces and active sites, and the activated basal plane of MoS2. The strong synergistic and coupling effects between 1T-MoS2 and Ir atoms markedly accelerated the reaction kinetics in alkaline solution.

Composite designing is the most attractive strategy for obtaining layered TMS-based bifunctional electrocatalysts for overall water splitting, and it involves integrating layered TMS with other active materials to construct novel heterostructures or composites. As we know, layered TMS can be employed as excellent HER electrocatalysts while they exhibit poor electrocatalytic OER performance. In turn, first-row metal compounds generally are preferred alternatives to noble metal oxides (IrO2, RhO2, or RuO2) in the OER, such as metal sulfides, oxides, and hydroxides.24 Hence, assembling a layered TMS with high activity for the HER with an active OER electrocatalyst in one system is a good choice for building highly effective bifunctional electrocatalysts. The newly formed composite/heterostructure would possess enhanced electrochemical water-splitting performance due to the synergistic effect of the counterparts.

Based on this design concept, Zhang et al. fabricated MoS2/Ni3S2 heterostructures with abundant interfaces on nickel (Ni) foam via decorating MoS2 nanosheets on the surfaces of Ni3S2 nanoparticles (Fig. 12a–c).93 The novel MoS2/Ni3S2 heterostructures possessed both highly efficient OER and HER activities in alkaline solution, with overpotentials at 10 mA cm−2 as low as 218 mV and 110 mV, respectively. Furthermore, an alkaline electrolyzer assembled using the as-prepared MoS2/Ni3S2 heterostructures could deliver a current density of 10 mA cm−2 at a low cell voltage of 1.56 V. Based on DFT calculations, the constructed interfaces between MoS2 and Ni3S2 could synergistically facilitate the chemisorption of hydrogen- and oxygen-containing intermediates, thus enhancing the electrocatalytic performance for water splitting. Subsequently, Peng et al. prepared three-dimensional (3D) Co9S8/WS2 array films on titanium (Ti) foil, in which WS2 nanosheets were uniformly grown in situ on the surfaces of Co9S8 array structures to form hybrids with large electrochemically active surface areas.94 Co9S8/WS2 array films displayed superior electrocatalytic activity for both the HER and OER with a low cell voltage of 1.65 V at 10 mA cm−2 when acting as both the cathode and the anode. The good performance for water splitting was ascribed to the unique porous structure and the synergistic effect of electronic and chemical coupling between active WS2 and Co9S8, which resulted in suitable kinetics and energetics for hydrogen adsorption and H2 formation processes and facilitated electron transport during water electrolysis. To further strengthen the activity and durability, the construction of oxygenated-sulfide hetero-nanosheets was subsequently put forward. Hou et al. grew vertically aligned oxygenated-MS2–MoS2 (M = Fe, Co, or Ni) hetero-nanosheets on carbon fiber cloth using bimetal precursors through a hydrothermal strategy.95 Among these samples, O-CoS2–MoS2 hetero-nanosheets presented the best performance, with low overpotentials of 97 and 272 mV to deliver a current density of 10 mA cm−2 for the HER and OER, respectively, in alkaline electrolyte. This unique architecture with well-exposed active heterointerfaces, ultra-fast charge-transport channels, and abundant active sites was highly favorable for obtaining enhanced electrocatalytic performance, which was derived from the synergistic effect of heterostructures. When assembled in a two-electrode electrolyzer, O-CoS2–MoS2 hetero-nanosheets merely needed a low cell voltage of 1.6 V to reach a current density of 10 mA cm−2.


image file: d0ta12152e-f12.tif
Fig. 12 (a) The proposed HER and OER mechanisms on MoS2/Ni3S2 heterostructures; S: yellow, Mo: blue, Ni: green, H: white, and O: red. (b) Polarization curves and (c) CV curves from MoS2/Ni3S2 heterostructures, MoS2 nanosheets, Ni3S2 nanoparticles, Ni foam, and IrO2/Pt.93 (d) A schematic diagram of interfacial electron transfer between MoS2 and Co3O4. (e) A free energy diagram of the HER on Co3O4@MoS2 and MoS2. (f) A free energy diagram of the OER on Co3O4@MoS2 and Co3O4. (g) Polarization curves of MoS2, Co3O4, and Co3O4@MoS2 in 1 M KOH.96

Motivated by this research progress relating to active bimetal transition-metal sulfides, Yang et al. proposed the construction of 3D composites co-assembled from MoS2 and Co9S8 nanosheets that were attached to Ni3S2 nanorod arrays.97 The as-prepared MoS2/Co9S8/Ni3S2 supported by Ni foam showed controllable morphologies and compositions, and it demonstrated excellent bifunctional electrocatalytic performance for both the HER and OER over a full pH range. Notably, the optimal sample could reach 10 mA cm−2 at very low overpotentials of 103, 117, and 113 mV for the HER in alkaline, neutral, and acidic solutions, respectively, while it required extremely low overpotentials of 166, 228, and 405 mV for the OER. In addition, a two-electrode electrolyzer assembled from MoS2/Co9S8/Ni3S2 composites also only required low cell voltages of 1.54, 1.80, and 1.45 V to obtain a current density of 10 mA cm−2 in alkaline, neutral, and acidic media, respectively. The superior performance of the all-pH bifunctional MoS2/Co9S8/Ni3S2 electrocatalyst was attributed to enhanced electron and mass transfer, and the well-exposed active sites of MoS2 and Co9S8 supported by a substrate with a 3D frame. The development of universal-pH bifunctional electrocatalysts has great significance for industrial applications involving seawater electrolysis.

As well as metal sulfides, layered TMS can also be integrated with other synergistic materials into well-designed architectures to achieve highly effective bifunctional electrocatalysts, like metal oxides, metal carbonitrides, and metal oxyhydroxides. Wang et al. innovatively prepared a WS2 nanosheet array on carbon cloth and then decorated NiO@Ni onto the surface via electrodeposition and subsequent thermal oxidation.98 This NiO@Ni/WS2 synergistic electrocatalyst possessed Pt-like activity for the HER with only a low overpotential (40 mV) being needed to drive 10 mA cm−2. Moreover, when serving as an anode for the OER, it required just 347 mV to deliver a current density of 50 mA cm−2, which was superior to even commercial RuO2. Excitingly, an electrolyzer with a two-electrode setup based on NiO@Ni/WS2 composites could afford 10 mA cm−2 at an extremely low cell voltage of 1.42 V. The marvelous enhanced bifunctional electrocatalytic performance of the NiO@Ni/WS2 composite may arise from the unique nanosheet array structure and multicomponent synergetic effects among WS2, NiO, and Ni. What is more, Liu et al. designed Co3O4@MoS2 heterostructures with balanced HER and OER performances.96 Theoretical and experimental results revealed that the heterostructures synergistically favored reducing the energy barrier for initial water dissociation, optimizing subsequent H adsorption/desorption for the HER, and enhancing the adsorption of oxygen intermediates for the OER (Fig. 12d–g). Analogously, Ji et al. grafted thin MoS2 nanosheets onto MOF-derived porous Co–N–C flakes and grew them on carbon nanofibers to get excellent bifunctional CoNC@MoS2/CNF films.99 The resultant CoNC@MoS2/CNFs with hierarchical structures and superior flexibility exhibited remarkable activities and stabilities for both the HER and OER in 1 M KOH. Regarding water electrolysis, a low cell voltage of 1.62 V was required to generate a current density of 10 mA cm−2, which could be ascribed to the high activity and rapid mass transport derived from a synergistic effect between MoS2 and Co–N–C flakes. Undoubtedly, MoS2 as a remarkable and representative layered TMS should be widely investigated to exploit its capacity for water splitting.

The stable phase of MoS2 is the thermally preferred 2H phase, while free-standing 1T-MoS2 suffers from an unstable nature. Integrating MoS2 with other active materials can not only improve the electrocatalytic activity significantly but can also stabilize unstable 1T-MoS2via serving as a substrate. Shang et al. embedded atomically thin 1T-MoS2 in quasi-amorphous CoOOH via a facile one-pot method to achieve a bifunctional electrocatalyst for overall water splitting.100 The synthesized 1T MoS2–CoOOH heterostructure exhibited highly efficient activity, with overpotentials of 158 mV in 0.5 M H2SO4 for the HER and 345 mV in 1 M KOH for the OER at 10 mA cm−2. In this heterostructure, the vertical 1T-MoS2 monolayer contributed mainly to the excellent HER activity, while the CoOOH nanosheets could provide superior OER activity. Importantly, the CoOOH nanosheets could also anchor and stabilize the 1T-MoS2 monolayer, achieving a stable architecture, via acting as an electron donor to facilitate the phase transition and stabilization of the metallic phase, which also paved the way to the controllable synthesis of thermodynamically unstable phases of layered TMS.

Heteroatom doping is an effective strategy for regulating the fundamental properties of layered TMS and constructing efficient bifunctional electrocatalysts. On one hand, the heteroatoms can directly break the periodic crystalline structure of materials via generating defects and distortions in the lattices of layered TMS; this is mainly due to differences between the atomic radii of host atoms and heteroatoms. These defects and distortions can create more active sites for electrocatalytic reactions and increase the reactivity of the surface.101 On the other hand, the introduction of heteroatoms can modulate the surface electronic structure of layered TMS through breaking the original charge balance, building a new charge distribution, and generating localized charge aggregation. The modulated electronic structure could further optimize the chemical adsorption energies of reactive intermediates and change the reaction pathways during water splitting.102 All these advantageous effects can promote the formation of electrocatalysts with bifunctionality and enhanced performance for water splitting. Heteroatom doping strategies can be divided mainly into two types: substitutional doping and surface charge transfer doping.103

In the case of substitutional doping, metal atoms or S atoms in layered structures are substituted with heteroatoms with different valence electrons. Xiong et al. developed a covalent Co doping strategy to induce bifunctionality into MoS2 for effective water splitting (Fig. 13c).104 Co atoms replaced Mo atoms to obtain MoS2 covalently doped with Co, which possesses enhanced intrinsic conductivity, a decreased ΔGH value upon optimizing the electronic structure of MoS2, and improved OER activity, which was ascribed to Co species in a high valence state under anodic potentials. Consequently, the best-performing MoS2 covalently doped with Co exhibited overpotentials of 48 and 260 mV to afford a current density of 10 mA cm−2 during the HER and OER, respectively, in 1.0 M KOH. A resultant electrolyzer assembled from MoS2 covalently doped with Co could generate 4.5 and 9.1 μmol min−1 O2 and H2, respectively, with nearly 100% faradaic efficiency. Specially, it should also be noted that the covalent doping of Ni and P into 1T-enriched MoS2 could significantly activate the basal planes, expand the interlayer spacing, and increase the edge density of 1T-enriched MoS2 in bifunctional nanostructures for water splitting.105 Further, Xue et al. prepared bifunctional Fe-doped MoS2 nanocanopies on Ni foam (Fe-MoS2/NF) via an in situ solvothermal reaction.106 Fe doping would modify the electronic structure of Mo, and some Fe would likely be present as iron sulfide. Strong coupling interactions between Fe and MoS2 markedly promote the electrocatalytic activity. The Fe-MoS2/NF electrode exhibited an OER overpotential of 230 mV at 20 mA cm−2 and a HER overpotential of 153 mV at 10 mA cm−2 in 1.0 M KOH.


image file: d0ta12152e-f13.tif
Fig. 13 (a) HER activity as a function of the Gibbs free adsorption energy of hydrogen and the configuration of TM/T1-vacancies. (b) The OER overpotentials of TM-doped T1 vacancy terminations; TM is a transitional metal and a T1 vacancy means that 37.5% of S atoms are added to initial Mo termini.107 (c) Side and top views of the optimized structure of MoS2 covalently doped with Co; Mo: green, S: yellow, and Co: blue.104 (d) The optimized geometries of CoRu(002)–MoS2 with absorbed intermediates; Mo: grayish purple, S: yellow, Ru: purple, Co: green, H: cyan, and O: red. (e) Gibbs free energy diagrams of the H2 production pathway (via the Volmer–Heyrovsky reaction in alkaline solution) on different samples.108

Surface charge transfer doping is realized via electron exchange between layered TMS and materials adsorbed on their surfaces. The introduction of doped heteroatoms is a viable method for triggering the bifunctional electrocatalytic activity of layered TMS, as has been proved via theoretical methods. Cheng et al. screened 28 transition-metal single-atom catalysts (SACs) supported on the edges of MoS2 as bifunctional electrocatalysts based on DFT calculations (Fig. 13a and b).107 These SACs could trigger OER activity and maintain the HER performance of MoS2 edges. The OER activity was calculated via a simple equation including the electronegativity of the atoms in the local structure and the coordination number of active metal centers. The lowest theoretical overpotentials for the HER and OER were derived from MoS2 edges modified with Pt single atoms. Pt single atoms enhanced the intrinsic HER activity of MoS2via modifying the bonds between S/Mo atoms and H and improved the OER activity via influencing the coordination environment and adsorption energy. This work provided theoretical guidance for developing innovative bifunctional electrocatalysts for overall water splitting.

Additionally, the synergistic effect of substitutional doping and surface charge transfer doping could lead to the great enhancement of the electrocatalytic performance of layered TMS. Kwon et al. doped metallic 1T′-MoS2 with Co atoms and further decorated this material with Ru nanoparticles (size = 1–5 nm) to prepare unique Co–Ru–MoS2 nanosheets for overall water splitting (Fig. 13d and e).108 Co atoms substituted for Mo atoms homogeneously and Ru nanoparticles were deposited on the surface of Co-doped MoS2. The enhanced HER performance in 1.0 M KOH was characterized based on the low overpotential of 52 mV at 10 mA cm−2 and small Tafel slope of 55 mV dec−1, and this was attributed to favorable water adsorption and dissociation on Ru atoms and advantageous reaction intermediates being catalyzed by Co atoms. The high OER performance benefitted from active RuO2 that was generated due to the oxidation of Ru nanoparticles, exhibiting a low overpotential of 308 mV at 10 mA cm−2 and a small Tafel slope of 50 mV dec−1. In short, the synergistic effect of doped Co atoms and active Ru nanoparticles could bring about excellent bifunctional electrocatalytic performance for overall water splitting via producing optimized electronic structures.

In conclusion, layered TMS exhibit excellent intrinsic electrocatalytic HER activities due to their well-exposed edge sites and good electrical conductivity, while their bifunctional electrocatalytic performances for water electrolysis could be optimized via some strategies such as phase engineering, composite designing, and heteroatom doping. Currently, MoS2-based and WS2-based bifunctional electrocatalysts are the most widely investigated materials in the layered TMS family, and they were always coupled with other active materials. To reach a current density of 10 mA cm−2 in the HER, a NiO@Ni/WS2 composite only requires an overpotential of 40 mV with a low Tafel slope of 83.1 mV dec−1 in 1.0 M KOH, which is even superior to Pt.98 To obtain 10 mA cm−2 in the OER, the best performance is shown by a MoS2/Co9S8/Ni3S2 composite, which only needs a small overpotential of 116 mV in 1.0 M KOH.97 A summary of current layered TMS-based bifunctional electrocatalysts is shown in Table 1, and further investigations should put more effort into the design of novel composites.

Table 1 A summary of layered TMS-based bifunctional electrocatalysts for overall water splitting
TMS-based material Preparation method Electrolyte Reaction/J (mA cm−2) Overpotential η (mV) Tafel slope (mV dec−1) Cell voltage (V @ mA cm−2) Stability
2H-TaS2 nanoflakes89 Solid-state + hydrothermal 1 M KOH HER/10, OER/10 145, 315 121, 81
Co9S8/WS2 array films94 Hydrothermal + solid-phase synthesis 1 M KOH HER/10, OER/10 138, — 80.2, — 1.65 24 h
NiO@Ni/WS2 (ref. 98) Hydrothermal + solid-phase synthesis 1 M KOH HER/10, OER/50 40, 347 83.1, 108.9 1.42 40 h
MoS2/Ni3S2 heterostructures93 Solvothermal 1 M KOH HER/10, OER/10 110, 218 83, 88 1.56 10 h
MoS2/Ni3S2 heteronanorods161 Hydrothermal 1 M KOH HER/10, OER/10 98, 248 61, 57 1.50 48 h
MoS2/Ni3S2 nanorod arrays144 Hydrothermal 1 M KOH HER/10, OER/10 187, 217 90, 38 1.467 24 h
MoS2/Ni3S2/Ni foam162 Electrodeposition + solvothermal 1 M KOH HER/10, OER/10 99, 185 71, 46 1.50 48 h
MoS2/NiS2 nanosheets151 Hydrothermal + solid-phase synthesis 1 M KOH HER/10, OER/10 62, 278 50.1, 91.7 1.59 24 h
MoS2/NiS yolk–shell microspheres163 Hydrothermal 1 M KOH HER/10, OER/10 244, 350 97, 108 1.64 24 h
Ni3S2–MoSx/Ni foam164 Solvothermal 1 M KOH HER/10, OER/50 65, 312 81, 103 1.54 20 h
CoNC@MoS2/CNF films99 Carbonization + solvothermal 1 M KOH HER/10, OER/10 143, 350 68, 51.9 1.62 56 h
MoS2/Co9S8/Ni3S2 composites97 Hydrothermal 1 M KOH HER/10, OER/10 113, 116 85, 58 1.54 24 h
0.5 M H2SO4 HER/10, OER/10 103, 228 55, 78 1.45 80 min
1.0 M PBS HER/10, OER/10 117, 405 56, 71 1.80 20 h
MoS2/Co9S8/N-carbon heterostructures165 Solvothermal 1 M KOH HER/10, OER/10 95, 230 94, 77 1.63 25 h
1T MoS2–CoOOH100 Hydrothermal 0.5 M H2SO4, 1 M KOH HER/10, OER/10 158, 345 42, 59 20 h
O–CoS2–MoS2 heteronanosheets95 Hydrothermal 1 M KOH HER/10, OER/10 97, 272 70, 45 1.60 10 h
CoSx@MoS2 heterostructures166 Solvothermal 1 M KOH HER/10, OER/10 146, 276 56.5, 26.1 1.668 20 h
Ni–Fe-LDH/MoS2 superlattices167 Exfoliation + assembly 1 M KOH HER/10, OER/10 180, 250 67, 45 1.4 11 h
(Ni, Fe)S2@MoS2 heterostructures168 Hydrothermal 1 M KOH HER/10, OER/10 130, 270 101.22, 43.21 1.56 45 h
1T′-MoS2/(Co, Fe, Ni)9S8 nanotube arrays118 Solvothermal 1 M KOH HER/10, OER/10 58, 184 37.5, 49.9 1.429 80 h
MoS2–NiS2/N-doped graphene169 CVD 1 M KOH HER/10, OER/10 172, 370 70, — 1.64 24 h
CoS–Co(OH)2@MoS2+x (ref. 170) Hybridization 1 M KOH HER/10, OER/10 143, 380 68, 68 1.58 11 h
CoS2–C@MoS2 nanofibers171 Hydrothermal 1 M KOH HER/10, OER/10 173, 391 61, 46 1000 cycles
CoS2–MoS2 hollow spheres172 Solvothermal 1 M KOH HER/10, OER/10 109, 288 52, 62.1 1.61 10 h
Co3S4@MoS2 heterostructures173 Hydrothermal 1 M KOH HER/10, OER/10 136, 280 43, 74 1.58 10 h
Co9S8@MoS2/carbon nanofibers120 Solid-phase synthesis 1 M KOH HER/10, OER/10 190, 430 110, 61 12 h
Co3O4/MoS2 heterostructures96 Hydrothermal 1 M KOH HER/10, OER/10 90, 269 59.5, 58 1.59 12 h
Co3O4/MoS2 heterostructures174 Hydrothermal 1 M KOH HER/10, OER/20 205, 230 98, 45 14 h
Ni2P/MoO2@MoS2 nanomaterials175 Solid-phase synthesis + phosphorization 1 M KOH HER/10, OER/10 159, 280 77, 85 1.72 40 h
Phosphorene quantum dot/MoS2 nanosheets176 Electrochemical 0.1 M KOH HER/10, OER/10 600, 370 162, 46 15 h
N–Ni3S2/N–MoS2 heteronanowires177 Solid-phase synthesis 1 M KOH HER/10, OER/200 54, 387 104.5, 101.2 1.79 @ 100 15 h
MoS2/NiS nanosheets178 Solvothermal + solid-phase synthesis 1 M KOH HER/10, OER/15 92, 271 113, 53 1.61 80 h
Co3S4@MoS2–Ni3S2 nanorods179 Hydrothermal 1 M KOH HER/10, OER/50 136, 270 72, 69 30 h
MoS2/NiCoS nanosheets180 Solvothermal 1 M KOH HER/10, OER/10 189, 290 75, 77 1.5 22 h
MoS2/Fe5Ni4S8 heterostructures181 CVD 1 M KOH HER/10, OER/10 120, 204 45.1, 28.1 10 h
Amorphous Ni–Co complexes/1T-MoS2 (ref. 182) Solvothermal 1 M KOH HER/10, OER/10 70, 235 38.1, 45.7 1.44 48
MoS2/(Co, Ni and Fe) sulfide183 Solvothermal 1 M KOH HER/100, OER/10 359, 178 76, 35.7 32 h
Fe–MoS2/Ni3S2 (ref. 184) Solvothermal 1 M KOH HER/10, OER/10 130.6, 256 112.7, 59.5 1.61 180 h
Co–MoS2 (ref. 104) Hydrothermal 1 M KOH HER/10, OER/10 48, 260 52, 85 11 h
Fe–MoS2 nanocanopies106 Solvothermal 1 M KOH HER/10, OER/20 153, 230 85.6, 78.7 1.52 140 h
Co–Ru–MoS2 nanosheets108 Hydrothermal 1 M KOH HER/10, OER/10 52, 308 55, 50 1.67 @ 20 16 h
Ir/MoS2 nanosheets92 Hydrothermal 1 M KOH HER/10, OER/10 44, 330 32, 44 1.57 12 h
Co–MoS2 nanosheets69 Hydrothermal 1 M KOH HER/10, OER/10 90, 190 50.3, 64.7 1.58
0.5 M H2SO4 HER/10, OER/10 60, 540 64.7, 245.9 1.90 ∼11 h


4.2 Bifunctional non-layered TMS (MxSy: Fe, Co, Ni, etc.) electrocatalysts

Non-layered TMS electrocatalysts are the most extensively studied bifunctional electrocatalysts for overall water splitting and they include FeS2, Ni3S2, and CoS2. Composed of first-row transition metals and S elements, these kinds of materials are very widely available, inexpensive, and chemically stable, making them quite valuable for energy conversion applications.80 Such bifunctional electrocatalysts based on Fe, Co, and Ni have been widely reported, frequently with similar or, in some cases, even better activities for both the HER and OER compared with electrocatalysts based on noble metals (like Pt, Ir, and Ru). To further improve their electrocatalytic performances for overall water splitting, many strategies have been developed, such as synergetic composite designing, heteroatom doing, strain engineering, facet engineering, edge engineering, and defect engineering.
4.2.1 Intrinsic bifunctional electrocatalysts. Ni-based sulfides are highly efficient non-noble metal-based bifunctional electrocatalysts with high conductivity and unique configurations, including Ni3S2, Ni3S4, and NiS.109 The controllable preparation of NiSx compounds with different phases and compositions can be easily realized via adjusting the added precursor ratios. Shi et al. prepared three NiSx compounds with large surface areas via a hydrothermal method, namely hexagonal NiS and cubic Ni3S4 and NiS2.110 Upon increasing the S/Ni molar ratio in the final product, the particle size decreases and the surface of the NiSx compound becomes rougher. A pyrite-type NiS2 electrocatalyst with enriched Ni3+ sites on the surface exhibited the best electrocatalytic activity, which was characterized by low overpotentials of 241 and 147 mV at 10 mA cm−2 for the OER and HER, respectively, in 1.0 M KOH. In practice, the enhanced performance could be attributed to the high-valence-state Ni3+ sites, which could optimize the binding energies with H* and OH* species, reduce the reaction barriers, and consequently speed up the sluggish kinetics. Similarly, Zheng et al. synthesized a series of NiSx nanocrystals with controllable phases and compositions via a facile polyol solution process, namely NiS, Ni3S2, and NiS2.111 Ni3S2 exhibited superior OER and HER activities compared to NiS and NiS2 in alkaline media, requiring overpotentials of 295 mV and 112 mV to afford 10 mA cm−2 for the OER and HER, respectively. The higher proportion of Ni in Ni3S2 leads to intrinsic metallic conductivity, abundant active sites, and optimized ΔGH values for the HER, additionally generating more active NiOOH species on the surface of Ni3S2 for the OER.

Obviously, Ni3S2 shows superior electrocatalytic performance owing to the high Ni element content, which could bring about intrinsic metallic behavior and a continuous network of Ni–Ni bonds throughout its whole structure. Zeng et al. reported 3D hierarchical Ni3S2 superstructures grown in situ on Ni foam via a chemical etching method.112 The porous rod-like array structure of Ni3S2 could provide large active surface areas, porosity, and highly effective accessibility. This desirable self-supported electrode offered ∼100% Faradaic yields for both the HER and OER, and it exhibited remarkable electrocatalytic stability for more than 50 h in 1.0 M KOH. Coincidentally, Zhang et al. fabricated 3D coral-like Ni3S2 on Ni foam (Ni3S2/NF) via a one-step electrochemical method.113 The optimal Ni3S2/NF electrode exhibited high electrocatalytic activity for overall water splitting, with low overpotentials of 89 and 242 mV for the HER and OER, respectively, to afford 10 mA cm−2. When the as-prepared Ni3S2/NF electrodes were assembled in an electrolyzer, a low cell voltage of 1.577 V was required to get 10 mA cm−2, with extremely long-term durability.

Ni-based sulfides can use commercial Ni foam directly as both a raw material and a growth template; the in situ growth of active Ni-based sulfides on Ni foam could further result in a self-supporting electrode for overall water splitting. Zhu et al. reported NiS microsphere film grown in situ on Ni foam (NiS/NF) via a vapor sulfurization reaction.114 This NiS/NF electrode could deliver 20 mA cm−2 with an overpotential of 158 mV for the HER and 50 mA cm−2 with an overpotential of 335 mV for the OER in 1.0 M KOH, exhibiting superior electrocatalytic activity and durability. The in situ strongly coupled interactions between NiS and Ni contributed to the enhanced electron interactions and lowered charge transfer barrier at the interface, which were beneficial for efficient electrocatalytic reactions. In fact, the interface between active Ni-based sulfides and Ni foam played a key role in generating their excellent electrocatalytic performance because a rational interface could create new electrochemically active sites via inducing unsaturated bonding, lattice distortion, and interfacial polarization.

Co-based sulfides also exhibit different phases and compositions, which can greatly affect the electrocatalytic activity and stability toward overall water splitting, with examples such as Co9S8, Co3S4, and CoS2. Ma et al. synthesized Co9S8, Co3S4, and CoS2 hollow nanospheres via adjusting the molar ratios of S and Co precursors using a facile solution-based strategy.115 The as-prepared CoS2 with a Co-rich component was superior to Co9S8 and Co3S4 in 1.0 M KOH, exhibiting overpotentials of 193 mV for the HER and 290 mV for the OER at 10 mA cm−2, and respective Tafel slopes of 57 and 100 mV dec−1. The electrocatalytic activities of these binary CoSx species were determined based on the coordination environments of Con+ active sites, the surface morphology, and the crystallographic structure. As is commonly known, metal atoms of transitional metal sulfides play important roles in determining electrocatalytic performance. Simultaneously, sulfur vacancies are also crucial in electrocatalytic processes. Water adsorption and dissociation on the low-index (100) facets of Co9S8 and Co3S4 have been investigated via a DFT approach.116 The sulfur vacancies on the (100) facet of Co9S8 enhanced the electrocatalytic activity toward water dissociation via elevating the energy level of unhybridized Co 3d states to the Fermi level, while having no significant impact on the energetics of the Co3S4 (100) facet.

Fe-based sulfides are emerging as a new generation of earth-abundant, low-cost, and highly active alternatives for water electrolysis. However, research into the OER performance of FeS2 is undertaken much less frequently than research into the HER electrocatalytic activity.38,81 Li et al. prepared FeS2/C nanoparticles on Ni foam via a facile hydrothermal strategy.117 These FeS2/C nanoparticles had excellent electrocatalytic activity, long-term durability, and fast charge-transfer kinetics, and they could deliver 10 mA cm−2 at overpotentials of 240 and 202 mV for the OER and HER, respectively.

4.2.2 Strategies for enhancing the electrocatalytic performance of non-layered TMS. Various inventive strategies have been developed for optimizing and improving the electrocatalytic performances of non-layered TMS via increasing the number of active sites and enhancing the intrinsic activity and electrical conductivity. Typically, synergetic composite designing, strain engineering, and heteroatom doping can significantly enhance the charge transfer kinetics during water electrolysis. Facet engineering, edge engineering, and defect engineering can improve the intrinsic electrocatalytic activities of non-layered TMS via exposing more active facets or creating more accessible active sites. In practice, these strategies are not mutually independent and sometimes can be applied simultaneously.

Designing composites is a conventional method for enhancing the charge transfer kinetics and creating new active sites; this can lead to a synergetic effect from each component and trigger the creation of additional interfaces and defects. Active species, like metal-based sulfides, oxides, and carbides, and carbon-based materials, all can be coupled with non-layered TMS to enhance bifunctional electrocatalytic activities. Li et al. reported complex polymetallic sulfide systems based on 1T′-MoS2 and (Co, Fe, Ni)9S8 as highly active bifunctional nanotube-array electrodes for water splitting, which were denoted as FeCoNi-HNTAs.118 The preparation process and microstructures of FeCoNi-HNTAs are shown in Fig. 14a–c. The electrode only required low overpotentials of 58 and 184 mV at 10 mA cm−2 for the HER and OER respectively, while maintaining good long-term stability. Further, when the electrodes were assembled into an electrolyzer, 10 mA cm−2 could be delivered at a low cell voltage of 1.429 V and excellent durability was exhibited (Fig. 14d and e). The outstanding electrocatalytic performance was principally thanks to the systematic optimization of the chemical composition and geometric structure. Abundant active sites, the excellent conductivity of metallic 1T′-MoS2, the synergistic effect of Fe, Co, and Ni ions, and the superaerophobicity of the electrode surface all contributed to the prominent electrocatalytic performance.


image file: d0ta12152e-f14.tif
Fig. 14 (a) Schematic diagrams of the preparation process of FeCoNi-HNTAs. (b) An FESEM image and (c) HRTEM image of the as-prepared FeCoNi-HNTAs. (d) Polarization curves and (e) the chronoamperometric response of FeCoNi-HNTAs in 1 M KOH with a typical two-electrode set-up.118 (f) FESEM, TEM, STEM, and STEM-EDS element mapping images of a Co9S8@MoS2 nanocrystal. (g) A HRTEM image of the Co9S8@MoS2 nanocrystal. Polarization curves of Co9S8@MoS2/CNFs, MoS2/CNFs, Co9S8/CNFs, and CNFs for (h) the HER in 0.5 M H2SO4 and (i) the OER in 1 M KOH.120

The interfaces between non-layered TMS and other materials in composites can bring about the proliferation of active sites. Li et al. constructed FeS2/CoS2 nanosheets with abundant defects at their interfaces as efficient bifunctional electrocatalysts.119 Benefiting from the disordered nanosheet interface structure with rich defects, FeS2/CoS2 nanosheets showed excellent electrocatalytic performance, with low overpotentials of 78.2 mV at 10 mA cm−2 for the HER and 302 mV at 100 mA cm−2 for the OER in 1.0 M KOH. More importantly, the FeS2/CoS2 nanosheets displayed remarkable performance for overall water splitting with a cell voltage of 1.47 V required to achieve 10 mA cm−2, and this activity could be maintained for more than 21 h. Zhu et al. designed and synthesized a hybrid system involving a cubic Co9S8@MoS2 core–shell structure on carbon nanofibers for overall water splitting, as shown in Fig. 14f–i.120 The Co9S8 nanoparticle core was surrounded by a shell of fullerene-like MoS2 layers to form a unique Co9S8@MoS2 core–shell structures on carbon nanofibers (CNFs) hybrid system. When compared with pure MoS2/CNFs and Co9S8/CNFs nanostructures, the hybrid system exhibited impressive enhanced HER and OER activities, which were attributed to a synergetic effect between Co and Mo in the proposed localized nano-interface region between Co9S8 and MoS2. The nano-interfaces could generate strong electron transfer between Mo and Co through intermediate S atoms bonded to both, which led to the promising improved electrocatalytic activity. Various composites based on non-layered TMS for overall water splitting have been exploited due to the diverse types of these materials, such as Ni3S2/Co9S8 arrays,121 NiS2/CoS2 nanostructures,122 interwoven NiS/NiS2 structures,123 and multiphase NiS–NiS2–Ni3S2.124

Metal-based oxides could also be applied as effective enhancers. Wu et al. reported ultrathin nanosheet-built, hollow MoOx/Ni3S2 microspheres on Ni foam that possess prominent electrocatalytic activity toward both the OER and HER.125 The unique structural features of the composite promoted the exposure of high-density electrocatalytically active sites, requiring an impressive low cell voltage of 1.45 V to reach 10 mA cm−2 with remarkable durability for more than 100 h in an alkaline electrolyzer. Furthermore, metal-based carbides, such as Mo2C with good electrical conductivity and a Pt-like d-orbital electronic structure, have been extensively studied for constructing composites with non-layered TMS, for example, porous Co9S8/N, S-doped carbon@Mo2C126 and a NiS2/carbon nanotubes@Mo2C bifunctional electrode.127

Carbon materials with different dimensions are another good choice for building composites based on non-layered TMS, like 0D graphene quantum dots, 1D carbon nanotubes, 2D graphene, and 3D graphite foam. Lv et al. fabricated nitrogen-doped graphene quantum dots (NGQDs) and Ni3S2 nanosheets on Ni foam (Ni3S2–NGQDs/NF).128 Self-supported Ni3S2–NGQDs/NF electrodes delivered 10 mA cm−2 at overpotentials of 216 and 218 mV for the OER and HER, respectively, and required a low voltage of 1.58 V at 10 mA cm−2 in an alkali electrolyzer. The experimental results and theoretical calculations proved that the excellent performance arose from synergistic effects between the constructed active interfaces of Ni3S2 and NGQDs; in other words, NGQDs could effectively regulate the electrocatalytic activity of Ni3S2. Analogously, Guo et al. synthesized metal sulfide (Ni3S2, Co9S8, or FeS) foams via the in situ conversion of corresponding metal (Ni, Co, or Fe) foams, and these then acted as templates for the surface growth of N-doped carbon nanotube (CNT) arrays for fabricating metal sulfide/CNT foam composites.129 Of these, Ni3S2/CNT foam served as the most effective electrode for overall water splitting with voltages of 1.5 and 1.72 V at 10 and 100 mA cm−2, respectively. Ni3S2 foam could provide abundant active sites and excellent conductivity; meanwhile, the CNT array coating could supply special superaerophobic geometry and protect the foam from electro-corrosion, indicating that tuning the surface geometry was an efficacious way to promote the electrocatalytic performance. Liu et al. grew vertically aligned ultrathin Co9S8 nanosheets in situ on N and S co-doped reduced graphene oxide (Co9S8/N,S-rGO) as an efficient electrocatalyst for water electrolysis.130 This unique hierarchical architecture was adopted to promote electron transport and fully expose active sites. Co9S8/N,S-rGO showed highly efficient and stable electrocatalytic performance for the OER and moderate HER activity in alkaline media, benefitting from a synergetic catalytic effect between Co9S8 nanosheets and N,S/rGO. Tong et al. prepared self-standing cobalt-disulfide/graphite foam (CoS2/GF) composite films via a hydrothermal approach.131 This 3D porous CoS2/GF electrode exhibited ultrahigh electrocatalytic activities for both the HER and OER, delivering a current density of 20 mA cm−2 at a cell voltage of 1.74 V in an alkaline electrolyzer. The excellent performances of CoS2/GF composite films were mainly attributed to the following factors: a strong synergistic effect between CoS2 and GF; the high electrical conductivity of GF, which could facilitate charge transfer during water electrolysis; the porous inter-layer structure of GF, which could provide effective pathways for releasing H2 and O2 bubbles; and the large surface area and numerous active sites of CoS2/GF composite electrodes.

Heteroatom doping plays a pivotal role in enhancing the electrocatalytic performances of non-layered TMS via regulating electronic structures to allow faster electron transfer and to optimize the binding energy of intermediates; strategies include substitutional doping and surface charge transfer doping. The substitutional doping of heteroatoms has attracted widespread attention in recent years, for example, the nitrogen doping of CoS2 could change the electronic density and minimize the adsorption free energy of materials.132 Yao et al. synthesized nitrogen-doped 3D dandelion-flower-like CoS2 structures directly grown on Ni foam (N-CoS2/NF).133 Benefiting from the unique 3D architecture and optimized hydrogen and water adsorption free energies due to N doping, an N-CoS2/NF electrode showed remarkable electrocatalytic performance, requiring only a low overpotential of 28 mV to gain 10 mA cm−2 for the HER and an overpotential of 200 mV at 20 mA cm−2 for the OER in alkaline solution (Fig. 15a and b). In fact, non-noble metal atoms are always applied as dopants due to the merits of high conductivity and low cost. Liu et al. reported pyrite NiS2 nanosheets doped with vanadium heteroatoms, with the conversion of the electronic structure from a typical semi-conducting form to one showing metallic characteristics.134 This interesting electronic structure reconfiguration of NiS2 was rooted in electron transfer from doped V sites to Ni sites, and it enabled Ni sites to obtain more electrons (Fig. 15e–g). Consequently, the metallic V-doped NiS2 nanosheets (with 10% V molar doping) displayed outstanding electrocatalytic performance, with overpotentials of 290 and 110 mV for the OER and HER, respectively, at 10 mA cm−2 in 1 M KOH.


image file: d0ta12152e-f15.tif
Fig. 15 (a) A top view of the optimized N-CoS2 (001) structure and (b) the d-orbital partial density of states of Co1 and Co2 from the areas labeled in (a).133 (c) The calculated water adsorption energy and (d) the HER free-energy change for N-Ni3S2 and pristine Ni3S2.135 (e) A schematic diagram of overall water splitting with V-NiS2 (with 10% V molar doping). (f) The calculated DOS of V-NiS2. (g) Charge-density contour maps of NiS2 and V-NiS2.134

Ternary and quaternary non-layered TMS with higher conductivity and richer redox reactions always exhibit greatly enhanced electrocatalytic performance, with examples such as Ni–Co–S, Fe–Co–S, and Ni–Fe–S systems.136 Sivanantham et al. directly grew NiCo2S4 nanowire arrays on the surface of 3D Ni foam (NiCo2S4 NW/NF) via a pressurized hydrothermal method.137 Benefiting from a high surface area, a well-distributed nanowire array structure, and enhanced charge transport due to the presence of mixed metals, this self-supported NiCo2S4 NW/NF electrode exhibited superior electrocatalytic performance, with low overpotentials of 260 and 210 mV at 10 mA cm−2 for the OER and HER, respectively, in 1 M KOH. Gong et al. synthesized FeCo2S4 nanosheets that were grown in situ on Ni foam (FeCo2S4/NF) using a facile hydrothermal sulfurization method.138 The self-supported FeCo2S4/NF electrodes delivered 10 mA cm−2, only requiring a small cell voltage of 1.541 V in a two-electrode alkaline electrolyzer. Furthermore, Yu et al. prepared hierarchical porous ternary nickel-iron sulfide (Ni0.7Fe0.3S2) microflowers on Ni foam via a traditional two-step method.139 Owing to the special 3D morphology and strong electron interactions between Ni, Fe, and S atoms, the as-synthesized electrocatalyst exhibited enhanced activity and durability for overall water splitting. The incorporation of Fe not only altered the morphology of sulfides from vertically aligned nanoplates to 3D hierarchical microflowers, but it also tuned the electronic structure of NiS2 and further optimized the binding energies of reactive intermediates. Furthermore, Li et al. reported hierarchical trimetallic sulfide FeCo2S4–NiCo2S4 nanosheet arrays supported on Ti mesh as efficient 3D bifunctional electrocatalysts for overall water splitting.140 This trimetallic sulfide electrode showed superior electrocatalytic performance, with low overpotentials of 230 and 150 mV at 10 mA cm−2 for the OER and HER, respectively, which could be attributed to the fully exposed active sites, fast charge transfer, and synergistic effects between multiple components.

Surface charge transfer doping with heteroatoms can be used to modify the electronic structures and morphologies of non-layered TMS simultaneously. Chen et al. developed 3D N-anion-decorated Ni3S2 on Ni foam (N-Ni3S2/NF) prepared via a one-step calcination route for both the OER and HER.135 Remarkably, the as-prepared N-Ni3S2/NF electrodes displayed extremely high electrocatalytic activity, long-term stability, and favorable reaction kinetics for both OER and HER processes. Low overpotentials of 330 and 110 mV were needed to reach 100 and 10 mA cm−2 for the OER and HER, respectively, in 1.0 M KOH, and a very low cell voltage of 1.48 V was required to deliver a current density of 10 mA cm−2 in an overall water-splitting device. The introduction of N anions notably modified the electronic structure and morphology of Ni3S2, leading to fully exposed active sites, enhanced electrical conductivity, and optimized hydrogen and water adsorption energies (Fig. 15c and d). Wang et al. synthesized metallic Co9S8 decorated with single-atomic Mo as an extremely efficient bifunctional electrocatalyst under various pH conditions.141 The superior electrocatalytic performance of Mo–Co9S8 could be attributed to the following factors: synergistic effects between atomically dispersed Mo and high oxidation state Co led to greatly increased activity and durability; the increased active surface area and superaerophobic interface could facilitate gas release; and the regulation of the electronic structure of Co9S8via the immobilization of single-site Mo species on its surface could enhance the electrical conductivity, with a higher density electron cloud and altered binding energies of adsorbed intermediate species.

Facet engineering is a unique method for increasing the intrinsic electrocatalytic activity of non-layered TMS via exposing more highly active facets. As is widely known, transition metals in non-layered TMS play a crucial role in determining the electrochemical performance, and they are always responsible for the good electrocatalytic activity. On one hand, transition metal cations on exposed low-index facets usually involve an unsaturated coordination environment, which would possess high electrochemical activity. For example, the low-index (100) facet of FeS2 is terminated with nonpolar [S–Fe–S] repeated along the regular facet direction, while Fe cations with five-fold coordination are located in a square pyramidal environment, as seen in Fig. 3d.38

On the other hand, the electrocatalytic activities of high-index facets with special structures, such as an asymmetric zigzag structure, will be higher than those of low-index facets in non-layered TMS. Feng et al. reported that the exposed (−210) high-index facet with an asymmetric zigzag structure in Ni3S2 nanosheet arrays contributed to the remarkable catalytic activity, which was much higher than the (001) low-index facet in Ni3S2.142 The as-prepared Ni3S2 nanosheet arrays with the (−210) high-index facet were synthesized via the direct sulfurization of Ni foam in a hydrothermal system and exhibited efficient and stable electrocatalytic activity toward both the HER and OER. DFT computations also elucidated that the HER and OER activities of the (−210) surface were better than those of the low-index (001) surface (Fig. 16a–c). Similarly, Li et al. synthesized ultrathin metallic CuFeS2 nanosheets with abundant exposed high-index (0−24) facets.143 Theoretical calculations and experimental results confirmed that the CuFeS2 nanosheets with exposed (0−24) high-index facets exhibited excellent electrocatalytic HER activity, benefitting from high electrical conductivity, optimized water adsorption energy, and the fast transformation efficiency of adsorbed H into H2 (Fig. 16d–f). However, these CuFeS2 nanosheets also have great potential for electrocatalyzing the OER.


image file: d0ta12152e-f16.tif
Fig. 16 (a) HRTEM and fast Fourier transform images of Ni3S2. (b) The most stable terminations of the (−210) and (001) surfaces of Ni3S2; the yellow and blue spheres represent S and Ni atoms, respectively. (c) A calculated free-energy diagram for the HER on the (−210) and (001) surfaces of Ni3S2 at equilibrium potential.142 (d) A HRTEM image of CuFeS2 nanosheets. (e) A chemisorption model of H on the (0−24) facet of CuFeS2 nanosheets; the brown, yellow, blue, and green spheres represent Cu, S, Fe, and H atoms, respectively. (f) A calculated free energy diagram for the HER using different materials.143

In brief, optimizing the density of crystallographic facets with a high fraction of low-coordinated atoms or unique asymmetric zigzag structures could be an effective way to improve the electrocatalytic activities of non-layered TMS. However, facet engineering strongly depends on the synthetic methods used for the controllable exposure of highly active TMS facets.

Edge engineering is a universal route for remarkably improving the electrocatalytic activity of non-layered TMS via creating more active sites or increasing their number; it is always presented as a change in structure or morphology. Regular arrangements and the design of unique structures of active materials are conducive to the full exposure of active sites on edges. Zhang et al. fabricated 3D hierarchical MoS2/Ni3S2 nanorod arrays that were nearly vertically aligned on Ni foam (MoS2/Ni3S2/NF) via a one-step hydrothermal procedure.144 Thanks to the elaborately designed and well-aligned architecture, the MoS2/Ni3S2/NF electrode with a large surface area and abundant active sites owned excellent electrocatalytic activity and durability for water electrolysis in an alkaline electrolyte. The hierarchical nearly vertically aligned structure possessed enough space for the diffusion of electrolyte into the well-exposed active sites, thus notably increasing the effective electrochemically active surface area of the electrode.

Guan et al. prepared self-supported hollow CoS2 nanotube arrays with hierarchical pores and rich active sites on a flexible substrate, which were derived from uniform wire-like metal–organic framework (MOF) nanoarrays.145 The open pore space and numerous nanoscale grains of hollow CoS2 nanotube arrays could provide a large surface area, a high density of active sites, and enhanced gas releasing abilities. The well-defined connection of active CoS2 with the flexible support ensured high levels of electron transport and good electrode mechanical stability. Consequently, the unique hollow CoS2 nanotube arrays exhibited high electrochemical activity, with a cell voltage of 1.67 V required to deliver a current density of 10 mA cm−2.

It is worth mentioning that the nanocrystallization of TMS materials is also a simple and promising way to increase the number of active sites via edge engineering. Yan et al. synthesized mulberry-like NiS/Ni nanoparticles via a surface modification strategy.146 Unique nanostructures of NiS/Ni could provide more active edges sites. Therefore, NiS/Ni nanoparticles exhibited excellent electrocatalytic activity in alkaline solution with remarkably low Tafel slopes of 46 and 74 mV dec−1 and small overpotentials of 301 and 161 mV for the OER and HER, respectively. In fact, all the TMS nanostructures mentioned here demonstrated the importance of nanocrystallization.

Strain engineering offers a novel path for promoting the electrocatalytic performance of non-layered TMS for efficient water splitting via inducing more active sites.147 The effects of strain engineering on layered NbS2 and MoS2 monolayers have been proved by Chen et al.148 They considered a wide strain range covering both tensile (0–10%) and compressive (0–6%) regions. Biaxial tensile strain enhanced the HER activity of layered TMS monolayers more effectively than uniaxial tensile strain, while compressive strain deteriorated the HER activity. Theoretical calculations revealed that tensile strain could reduce the adiabatic proton affinity but simultaneously increase the adiabatic electron affinity to a larger extent, activate relatively inert inner valence electrons, and further enlarge the d-band exchange splitting, all of which could enhance the electrocatalytic HER activity of layered TMS.

Conceivably, this strategy is also suitable for non-layered TMS. Shan et al. predicted strain-induced changes in the electronic structure and electrocatalytic performance of a 2D Fe3S4 material.149 The calculation results revealed that the half-metallic features of Fe3S4 would be converted into metallic features due to the small lattice deformation in Fe3S4 monolayers induced by applied external strain, which could optimize the Gibbs adsorption free energy of hydrogen and improve the electron transfer capabilities (Fig. 17a–c). However, more efforts should be carried out to investigate the influence of strain engineering on the OER performance of TMS-based electrocatalysts.


image file: d0ta12152e-f17.tif
Fig. 17 The charge densities of an optimized (1 × 1) Fe3S4 monolayer (a) without external strain and (b) with 1.0% compressed strain. (c) The Gibbs free energy of hydrogen adsorbed on the surface of a Fe3S4 monolayer as a function of the external strain.149 (d) The crystal structure of CoS2 and models of surface single-layer atoms in CoS2, N-doped CoS2 (Ns-CoS2), and N-doped CoS2 with a S vacancy (Ns-Vs-CoS2). (e) The HER free energy diagram for Co sites. (f) The HER free energy diagram for Ns sites.150

The defect engineering of non-layered TMS is not naturally independent, and it is usually combined with the formation of heterogeneous interfaces,151 unique architecture design,152 and heteroatom doping.153 Defects always have a positive effect on the adsorption activation energy of reactants and the dissociation energy of products in overall water splitting, facilitating the generation and conversion of intermediates via regulating the electronic structure.154 Here, attention is focused on vacancy defects in the lattice structure, which can be induced via special techniques such as plasma etching.85 Theoretical calculations based on first-principles studies revealed the effects of single cobalt and sulfur vacancies on structural and electronic changes via making pyrite CoS2 samples with high densities of defects and inducing localized defect states in the gap range of the minority spin channel.155 Yang et al. developed highly porous Co–S films that were grown in situ on Ni foam, which had superior electrocatalytic activity and a robust nature for water electrolysis.156 Self-supported Co–S film with a S content level of 21.7 at% could drive a high current density of 500 mA cm−2 with low overpotentials of 368 and 155 mV for the OER and HER, respectively. The optimal S content in Co–S film could create more vacancy defects, which would give rise to the proliferation of electrocatalytically active sites or oxygen deficiencies during the OER, significantly enhancing the intrinsic activity and reaction kinetics for both the HER and OER. As reported, sulfur vacancies and strain could be used to finely tune the HER electrocatalytic activity of monolayer MoS2.41 Similarly, Zhang et al. proposed that sulfur vacancies and nitrogen dopants in CoS2 could synergistically activate and optimize the HER activity via enhancing the activities of neighboring Co sites with optimal hydrogen adsorption free energies, as shown in Fig. 17d–f.150

All in all, non-layered TMS are the most extensively studied bifunctional electrocatalysts for water electrolysis with the advantages of being highly active, chemically stable, widely available, and cheap. To further improve their bifunctional electrocatalytic performances, many strategies have been proposed for increasing the number of active sites and enhancing the intrinsic activity and electrical conductivity, such as composite design, heteroatom doping, strain engineering, facet engineering, edge engineering, and defect engineering. To now, studies of non-layered TMS bifunctional electrocatalysts have mainly focused on Ni-based and Co-based sulfides. A summary of non-layered TMS-based bifunctional electrocatalysts is shown in Table 2. N-CoS2/Ni foam requires a quite low overpotential of 28 mV to reach 10 mA cm−2 for the HER in 1.0 M KOH, exhibiting the best electrocatalytic HER activity.133 Moreover, to reach a current density of 10 mA cm−2 in the OER, carbon dots/NiCo2S4/Ni3S2 composites require the smallest overpotential of 116 mV in 1.0 M KOH.157 Surprisingly, these non-layered TMS bifunctional electrocatalysts possess the potential to surpass noble-metal oxides in overall water splitting, showing far-reaching significance for large-scale water electrolysis applications.

Table 2 Summary of non-layered TMS-based bifunctional electrocatalysts for overall water splitting
TMS-based materials Preparation method Electrolyte Reaction/J (mA cm−2) Overpotential η (mV) Tafel slope (mV dec−1) Cell voltage (V @ mA cm−2) Stability
FeS2/C nanoparticles117 Hydrothermal 1 M KOH HER/10, OER/10 202, 240 98, 92 1.72 5 h
FeS nanosheets185 Solvothermal 1 M KOH HER/100, OER/10 243, 238 77, 82.7 50 h
Ni–FeS2 (ref. 186) Electrodeposition 1 M KOH HER/10, OER/100 157, 255 112.5, 58.3 1.588 11 h
Co–FeS2 nanospheres187 Solvothermal 1 M KOH HER/10, OER/10 267, 324 58, 50 1.60 10 h
NiS2/FeS2/N-doped carbon nanorods188 Carbonization + solid-phase synthesis 1 M KOH HER/10, OER/20 172, 231 115.64, 44.29 1.58 80 h
FeS/NiS/Ni foam189 Deposition + solvothermal 1 M KOH HER/10, OER/10 144, 203 120, 39 1.618 10 h
CoS2 hollow nanospheres115 Hydrothermal 1 M KOH HER/10, OER/10 193, 290 100, 57 1.54 10 h
CoS2 nanosheets190 CVD 1 M KOH HER/10, OER/10 90, 220 48, 92 1.58 22 h
CoS2 nanotube arrays145 Solid-phase synthesis 1 M KOH HER/10, OER/10 193, 276 88, 81 1.67 20 h
CoSx freestanding sheets191 Liquid-phase growth 1 M KOH HER/10, OER/10 127, 288 117, 91 1.55 @ 20 48 h
CoS nanoflake array192 Hydrothermal 1 M KOH HER/50, OER/10 247, 310 114.8, 73.4 1.72 15 h
Co9S8 nanosheets193 Hydrothermal 0.5 M H2SO4, 1 M KOH HER/10, OER/10 178, 206 82, 87 8 h
Co3S4 nanosheets194 Solvothermal 1 M KOH HER/10, OER/10 270, 360 124.5, 84.7 1.63 5 h
Co–S films156 Electrodeposition 1 M KOH HER/100, OER/100 124, 322 65.5, 66.4 1.79 150 h
CoS2/graphite foam131 Hydrothermal 1 M KOH HER/20, OER/20 224, 298 148, 82.6 1.74 @ 20 18 h
Co3S4@carbon microflowers152 Hydrothermal 1 M KOH HER/10, OER/10 140, 250 103, 78 1.58 8 h
Mo–Co9S8 nanorod array195 Hydrothermal 1 M KOH HER/10, OER/50 139, 210 57, 123 1.50 20 h
Cu–Co9S8 nanorod array196 Solid-phase synthesis 1 M KOH HER/10, OER/50 62, 260 76.7, 132.4 1.49 25 h
Cr–Co9S8 nanoarrays197 Hydrothermal + solid-phase synthesis 1 M KOH HER/10, OER/100 120, 313 138.8, 10.1 1.45 12 h
Ni–Co9S8 nanoarrays198 Hydrothermal + solid-phase synthesis 1 M KOH HER/10, OER/100 90, 300 62.8, 105.1 1.45 15 h
Ni–Co3S4 nanowires199 Hydrothermal 1 M KOH HER/10, OER/100 199, 283 91, 65 1.70 40 h
F–CoS2/NF composites200 Hydrothermal 1 M KOH HER/10, OER/100 112, 1596 61, 104 100 h
N–CoS2/Ni foam133 Hydrothermal + heating 1 M KOH HER/10, OER/20 28, 200 42.6, 55 1.50 11 h
P–CoS2/carbon cloth201 Solvothermal 1 M KOH HER/10, OER/10 80, 250 98, 90 1.56 24 h
Co0.9S0.58P0.42 (ref. 202) Solid-phase synthesis + phosphorization 1 M KOH HER/10, OER/10 140, 266 78, 48 1.59 30 h
TiO2@Co9S8 core-branch arrays203 Solid-phase synthesis 1 M KOH HER/10, OER/10 139, 240 65, 55 1.56 30 h
P–CoMoS/carbon cloth204 Hydrothermal + phosphorization 1 M KOH HER/10, OER/10 66, 260 60.1, 72.2 1.54 24 h
CoS@few-layer graphene205 High-temperature treatment 1 M KOH HER/10, OER/10 118, 350 63, 70.9 1.77 10 h
CoS2@N-doped graphene206 Hydrothermal 1 M KOH HER/10, OER/10 204, 243 108, 51.8 1.53 12 h
N-, O-, S-tridoped carbon-encapsulated Co9S8 (ref. 207) Pyrolysis 1 M KOH HER/10, OER/10 320, 340 105, 68 1.60 10 h
Co9S8/N,S-rGO130 Polyol refluxing + solid-phase synthesis + calcination 1 M KOH HER/10, OER/10 332.4, 266 131.4, 75.5 20 h
S,N-CNTs/CoS2@Co nanoparticles208 Self-catalysis + solid-phase synthesis 1 M KOH HER/10, OER/10 112, 157 104.9, 76.1 1.633 20 h
Co9S8/N,S-doped carbon@Mo2C126 Pyrolysis 0.5 M H2SO4 HER/10, OER/10 74, — 69.3, —
1.0 M phosphate solution HER/10, OER/10 121, — 106.4, —
1 M KOH HER/10, OER/10 89, 293 86.7, 59.7 1.61 20 h
Co9S8/MnS/Ni foam209 Solvothermal 1 M KOH HER/100, OER/100 217, 298 48.2, 43.9 1.47 20 h
CoNi2S4@CoS2/Ni foam210 Hydrothermal + solid-phase synthesis 1 M KOH HER/10, OER/10 173, 259 51, 45 16 h
Co9S4P4 pentlandite159 Etching + CVD 1.0 M PBS HER/10, OER/25.9 87, 570 51, 106 1.67 24 h
Zn0.3Co2.7S3P nanoparticles211 Solvothermal + phosphorization 1 M KOH HER/10, OER/10 188, 261 76.1, 56.2 1.70 10 h
CuCo2S4 nanowire arrays212 Hydrothermal 1 M KOH HER/10, OER/100 65, 310 42, 49 1.65 @ 100 50 h
CuCo2S4 (ref. 213) Hydrothermal 1 M KOH HER/10, OER/20 158, 290 113, — 1.66 24 h
CuCo2S4@Ni(OH)2 nanorods214 Hydrothermal + electrodeposition 1 M KOH HER/10, OER/100 117, 250 170, 103.4 1.47 20 h
FeCo2S4 nanosheet arrays215 Hydrothermal 1 M KOH HER/10, OER/50 132, 270 164, 59 1.56 20 h
FeCo2S4 nanosheet arrays138 Hydrothermal + solid-phase synthesis 1 M KOH HER/10, OER/20 131, 259 52, 173 1.541 12 h
Fe–Co–S nanosheets216 Hydrothermal 1 M KOH HER/10, OER/10 68, 208 114, 37 1.44 80 h
NiCo2S4 nanowire array217 Wet-chemistry solid-phase synthesis 1 M KOH HER/50, OER/50 266, 300 130.5, 90.9 1.66 20 h
NiCo2S4 nanowire arrays137 Hydrothermal 1 M KOH HER/10, OER/10 210, 260 58.9, 40.1 1.63 50 h
NiCo2S4 nanoflakes218 Reverse microemulsion + solid-phase synthesis 1 M KOH HER/10, OER/100 169, 319 97.1, 53.3 1.61 70 h
NiCo2S4 nanowires219 Hydrothermal 1 M KOH HER/10, OER/40 191, 256 116.9, 67.7 1.59 12 h
Ni4.3Co4.7S8 (ref. 220) Hydrothermal 1 M KOH HER/10, OER/20 148, 133.8 90, 194.2 1.67 @ 50
(Ni0.33Co0.67)S2 nanowires/carbon cloth221 Hydrothermal 1 M KOH HER/100, OER/100 334, 216 127, 78 1.57 30 h
Ni–Mo–S nanowires222 Electroplating + solid-phase synthesis 1 M KOH HER/100, OER/100 290, 390 103, 75 2 @ 100 12 h
NiFeMoS nanorods223 Electrodeposition + hydrothermal 1 M KOH HER/10, OER/150 100, 280 121, 69 1.52 @ 100 10 h
Ni8V2(Mo3S4)11/Ni–Co foam224 Hydrothermal 1 M KOH HER/10, OER/10 129, 330 136, 183 25 h
NiCo/NiCo2S4@NiCo arrays225 Hydrothermal + electrodeposition 1 M KOH HER/10, OER/100 132, 294 58.2, 59.6 1.55 27 h
FeCo2S4–NiCo2S4 nanosheet arrays140 Hydrothermal 1 M KOH HER/10, OER/10 150, 230 38, 65 1.51 10 h
Cu@CoSx/copper foam226 Chemical deposition 1 M KOH HER/10, OER/10 134, 160 61, — 1.50 200 h
Ni–Fe–Co–S nanosheets227 Electrodeposition 1 M KOH HER/10, OER/10 106, 207 95, 63 1.54 10 h
Ni–Co–S nanosheet arrays228 Annealing + solid-phase synthesis 1 M KOH HER/10, OER/10 110, 270 56, 80 1.62 12 h
P–(Ni, Fe)3S2 nanosheet arrays229 Phosphorization + solid-phase synthesis 1 M KOH HER/10, OER/10 98, 196 88, 30 1.54 15 h
(Ni–Fe)Sx/NiFe(OH)y hollow microtubes/spheres230 Electrodeposition 1 M KOH HER/100, OER/100 124, 290 68, 58 1.46 50 h
Co(OH)2/Ni–Co–S nanotube arrays231 Hydrothermal 1 M KOH HER/100, OER/100 254, 340 88, 64 1.62 20 h
Co9S8/Ni3Se2/graphene232 Hydrothermal + solvothermal 1 M KOH HER/20, OER/10 170, — 83, — 1.62 10 h
Ni–Mo–S@Co3O4/carbon fibers233 Hydrothermal 1 M KOH HER/10, OER/10 85, 275 50, 109 1.57 50 h
N-carbon coated NiCo2S4 hollow nanotubes234 Solvothermal 1 M KOH HER/100, OER/100 295, 330 89.8, 86.8 1.60 15 h
rGO/(NixMnyCoz)3S4 nanocomposites235 Solid-phase synthesis 1 M KOH HER/10, OER/10 151, 320 52, 58 1.56 @ 20 24 h
Nanoporous (003)-Ni3S2 (ref. 50) Anodization + solid-phase synthesis 1 M NaOH HER/10, OER/10 135, 175 75.7, 101.2 1.611 30 h
(−210)-Ni3S2 nanosheet arrays142 Solid-phase synthesis 1 M KOH HER/10, OER/10 223, 260 —, — 200 h
NiS microsphere film114 Solid-phase synthesis 1 M KOH HER/20, OER/50 158, 335 83, 89 1.64 35 h
Ni3S2 superstructures112 Chemical etching 1 M KOH HER/10, OER/100 135, 320 75, 59 1.59 50 h
Ni3S2 nanocrystals111 Ambient-pressure polyol solution 1 M KOH HER/10, OER/10 112, 295 67, 52 1.63 108 h
Ni3S2 nanorods236 Hydrothermal 1 M KOH HER/10, OER/10 82, 310 73.8, 80.1 1.61 30 h
Ni3S2 superstructures/Ni foam237 Chemical etching 1 M KOH HER/10, OER/10 182, 340 89, 150 12 h
Ni3S2/Ni foam238 Annealing 1 M KOH HER/10, OER/10 131, 312 96, 111 1.68 14 h
Coral-like Ni3S2/Ni foam113 Electrodeposition 1 M KOH HER/10, OER/10 89, 242 85, 74 1.577 40 h
Ni3S2/Ni foam239 Solvothermal 1 M KOH HER/10, OER/10 123, 222 91, 60 1.59 20 h
NiS2 nanospheres110 Hydrothermal 1 M KOH HER/10, OER/10 147, 241 105, 65 1.66 40 h
NiS2 nanowires/carbon fiber paper240 Solid-phase synthesis 1 M KOH HER/10, OER/10 165, 246 134, 94.5 1.59 20 h
NiS2 microspheres and NiS microspheres67 Hydrothermal + annealing 1 M KOH HER/10, OER/10 148, 320 79, 59 1.58 12 h
Ni7S6 foam241 Annealing 0.5 M H2SO4, 0.1 M KOH HER/10, OER/10 70, 140 63, 119 —, 1.51 24 h
NiS/Ni coupled interface146 Solid-phase synthesis 1 M KOH HER/10, OER/10 161, 301 74, 46 18.6 h
Ni/NiS/N-doped mesoporous carbon242 Solid-phase synthesis 1 M KOH HER/10, OER/10 70, 337 45, 52 1.61 25 h
Ni(S0.5Se0.5)2 hollow/porous spheres243 Hydrothermal + selenization 1 M PBS HER/10, OER/100 124, 501 81, 94 1.87 12 h
N–Ni3S2 (ref. 135) Annealing 1 M KOH HER/10, OER/100 110, 330 —, 70 1.48 8 h
Sn–Ni3S2 nanosheets244 Hydrothermal 1 M KOH HER/100, OER/100 170, 270 55.6, 52.7 1.46 60 h
Mo–Ni3S2 nanorods153 Hydrothermal 1 M KOH HER/100, OER/100 278, 180 72.9, 45.5 1.53 15 h
Mo–Ni3S2 nanosheets245 Hydrothermal 1 M KOH HER/10, OER/10 212, 260 98, 85 1.67 13 h
Mo–NiS heterostructures246 Hydrothermal + solid-phase synthesis 1 M KOH HER/10, OER/50 147.6, 390 88.1, 185 1.92 18 h
Mo/Mn–NixSy/Ni foam247 Hydrothermal + solid-phase synthesis 1 M KOH HER/10, OER/50 144, 162 91, 110 1.49 24 h
Ni0.7Fe0.3S2 microflowers139 Hydrothermal 1 M KOH HER/10, OER/10 155, 198 109, 56 1.625 14 h
Fe-Ni3S2 nanowires248 Solvothermal 1 M KOH HER/10, OER/200 109, 223 49.5, 55.7 1.53 14 h
Fe-Ni3S2 nanosheet arrays8 Hydrothermal + ion exchange 1 M KOH HER/10, OER/10 47, 214 95, 42 1.54 10 h
Fe11.1%-Ni3S2/Ni foam249 Hydrothermal + wet-chemistry conversion 1 M KOH HER/50, OER/100 203, 252 89, 61.7 1.60 42 h
Co–Ni3S2 nanoarrays250 Hydrothermal 1 M KOH HER/10, OER/10 240, 120 145, 38.4 1.69 @ 50 10 h
V–NiS2 nanosheets134 Solid-phase synthesis 1 M KOH HER/10, OER/10 110, 290 90, 45 1.56 20 h
V–Ni3S2 nanorod array251 Hydrothermal 1 M KOH HER/10, OER/10 133, 148 70, 42 1.421 60 h
V–Ni3S2@NiO/Ni foam252 Hydrothermal 1 M KOH HER/10, OER/10 112, 170 69, 98.8 1.52 55 h
(−210)-Ni3S2 branch arrays253 Solid-phase synthesis 1 M KOH HER/10, OER/10 112, 220 69, — 1.58 10 h
MoOx/Ni3S2 microspheres125 Solid-phase synthesis + hydrothermal 1 M KOH HER/10, OER/10 106, 136 90, 50 1.45 200 h
Ni3S2/VO2 core/shell nanoarray254 Hydrothermal 1 M KOH HER/10, OER/10 100, 150 114, 47 1.42 15 h
Ni3N/Ni3S2 (ref. 255) Calcination 1 M KOH HER/100, OER/100 197, 404 58.8, 112 12 h
Ni3S2/MnS nanosheets256 Hydrothermal 1 M KOH HER/10, OER/10 116, 228 41, 46 1.54 50 h
Se–MnS/NiS heterojunctions257 Hydrothermal chemical-deposition 1 M KOH HER/10, OER/10 56, 211 55, 50 1.47 48 h
CdS/Ni3S2 nanosheet flowers258 Hydrothermal 1 M KOH HER/10, OER/10 121, 151 110, 174 12 h
NixCo3−xS4/Ni3S2/Ni foam259 Cation exchange 1 M KOH HER/10, OER/10 136, 160 107, 95 1.53 200 h
CoMoS4/Ni3S2 nanostructures158 Hydrothermal + solid-phase synthesis 1 M KOH HER/10, OER/10 158, 200 169, 63 1.568 10 h
N-NiMoO4/NiS2 nanowires/nanosheets260 Solid-phase synthesis 1 M KOH HER/10, OER/10 99, 283 74.2, 44.3 1.60 20 h
Carbon dots/NiCo2S4/Ni3S2 (ref. 157) Hydrothermal + cation exchange 1 M KOH HER/10, OER/10 127, 116 148, 99 1.50 12 h
Mo(1−x)WxS2/Ni3S2 heterostructures261 Hydrothermal 1 M KOH HER/10, OER/10 98, 285 92, 98 1.62 50 h
NiS–CoS nanorod arrays262 Electrodeposition + hydrothermal 1 M KOH HER/10, OER/10 102, 170 92, 71 1.47 12 h
NiS2/CoS2 microstructures122 Solid-phase synthesis 1 M KOH HER/10, OER/20 165, 310 72, 78 1.61 24 h
Co9S8/Ni3S2 nanowire arrays263 Hydrothermal + vulcanization 1 M KOH HER/10, OER/10 128, 227 97.6, 46.5 1.64 9 h
Co9S8@Ni(OH)2 nanorods264 Hydrothermal + electrodeposition 1 M KOH HER/10, OER/60 100, 250 143.5, 273.4 1.55 10 h
CoSx/Ni3S2 heterostructures265 Hydrothermal 1 M KOH HER/10, OER/20 204, 280 113.1, 92.7 1.572 30 h
Ni3S2/Co9S8 arrays121 Solid-phase synthesis 1 M KOH HER/100, OER/100 269, 340 98, 66 1.55 12 h
CoS/Ni3S2–FeS nanopetals266 Electrodeposition + solid-phase synthesis 1 M KOH HER/10, OER/10 75, 136 57, 51 24 h
FeS2/CoS2 interface nanosheets119 Hydrothermal + solid-phase synthesis 1 M KOH HER/10, OER/100 78.2, 302 44, 42 1.47 21 h
Cu2S–Ni3S2 array267 Hydrothermal 1 M KOH HER/10, OER/10 149, 329 75.89, 44.11 1.77 100 h
Ni3S2/VS4 nanohorn array268 Solvothermal 1 M KOH HER/10, OER/50 177, 317 139, 43 1.57 70 h
NiS/NiS2 (ref. 123) Thermal decomposition 1 M KOH HER/100, OER/100 248, 416 95.1, 156.5 1.62 36 h
NiS–NiS2–Ni3S2/Ni foam124 Solid-phase synthesis 1 M KOH HER/20, OER/20 137, 143 87, 80 1.46 14 h
Mn-doped Cu7.2S4@NiS2@NiS269 Hydrothermal + solid-phase synthesis 1 M KOH HER/10, OER/50 87, 252 43.7, 130.1 1.514 46 h
Mo–Ni3S2/NixPy hollow nanorods270 Solvothermal 1 M KOH HER/10, OER/50 109, 238 68.4, 60.6 1.46 72 h
Ni2P/Ni3S2 heteronanoflake arrays271 Hydrothermal + phosphorization 1 M KOH HER/10, OER/10 80, 210 65, 62 1.50 36 h
NiS/Ni2P heterostructures272 Hydrothermal + solid-phase synthesis + phosphorization 1 M KOH HER/20, OER/20 111, 265 78.1, 41.3 1.67 10 h
N-Ni3S2/VS2 nanosheets273 Hydrothermal 1 M KOH HER/10, OER/10 151, 227 107.5, 70.5 1.648 20 h
Ni(OH)2/Ni3S2 nanosheet arrays274 Co-precipitation + solid-phase synthesis 1 M KOH HER/20, OER/20 270, 211 129, 152.7 1.57 10 h
NiS/Ni2P@Ni(OH)2 (ref. 275) Electrodeposition 1 M KOH HER/10, OER/10 120, 219 71, 82 1.58 160 h
NiFe-LDH@Ni3S2 nanosheet arrays276 Hydrothermal 1 M KOH HER/20, OER/50 197, 230 99, 29 1.65 12 h
Se–(NiCo)Sx/(OH)x nanosheets277 Solvothermal 1 M KOH HER/10, OER/10 103, 155 87.3, 33.9 1.60 66 h
N-CNTs/NiS2@Mo2C127 Calcination + CVD 1 M KOH HER/10, OER/10 227, 320 114.6, 77.5 1.52 24 h
Ni3S2/N-CNTs129 Solid-phase synthesis 1 M KOH HER/50, OER/50 135, 380 58, 77 1.721 @ 100 50 h
Ni3S2@graphite foam@Ni foam278 Hydrothermal 1 M KOH HER/10, OER/10 99, 240 98.2, 62.4 1.63 30 h
N-doped graphene QDs/Ni3S2 nanosheets128 Hydrothermal 1 M KOH HER/10 218, 216 89, 95.5 1.58 17 h
OER/10
3D graphene–Au–Ni3S2 (ref. 279) Electrodeposition 1 M KOH HER/10, OER/91.2 140, 370 93, 148 1.63 19 h


5. Conclusions and outlook

In summary, we have outlined the research progress relating to bifunctional electrocatalysts based on transition metal sulfides for effective overall water splitting, showing their potential as candidates to replace noble-metal-based electrocatalysts. The fabrication of TMS electrocatalysts via ‘bottom-up’ and ‘top-down’ strategies has been summarized. Typically, exfoliation and ball milling approaches are the leading ‘top-down’ strategy processing methods, which are suitable for synthesizing nano-structured and micro-structured TMS on a large scale from bulk TMS materials. Additionally, hydrothermal, solvothermal, chemical vapor deposition, and solid-phase synthesis approaches are the main representatives of the ‘bottom-up’ strategy, showing great potential for the preparation of self-supported TMS electrodes when suitable reaction substrates are adopted. Via these controlled fabrication approaches, the rational tuning of well-defined TMS shapes, sizes, hierarchical structures, and defect states could be achieved, with great influence on the electronic structure and electrocatalytic performance of TMS. Importantly, the active sites of TMS electrocatalysts are nearly always metal atoms, while exposed S atoms and S vacancies also play a vital direct role in the HER via increasing the number of active sites and changing the coordination environments of adjacent metal atoms. Moreover, the indirect role of S atoms in the HER provides locations for hydrogen adhesion and separation when metal atoms act as the active sites. However, TMS tend to function as pre-catalysts in complex OER processes, with involvement in conversion to transition metal hydr(oxy)oxides.

Layered TMS (e.g., MoS2, WS2, and TaS2) exhibit excellent intrinsic electrocatalytic HER activity due to their well-exposed edge sites and good electrical conductivity, while their direct application in the OER is limited. To further explore the potential of layered TMS for use in water splitting, numerous efforts have been carried out to optimize their electrochemical properties via improving their intrinsic electrical conductivity and stability and enhancing their chemisorption capabilities for O-containing species. Phase engineering, composite design, and heteroatom doping are favorable strategies for achieving this goal. Phase engineering is a promising way to enhance the electrocatalytic OER activity of layered TMS, as the 1T phase usually exhibits higher OER activity than the 2H phase. Integrating layered TMS with other materials with highly active OER performance or good conductivity can be used to construct bifunctional electrocatalysts with novel composites/heterostructures utilizing synergistic effects relating to the counterparts. Lastly, heteroatom doping is favorable for modulating the electronic structures of layered TMS and optimizing the chemical-adsorption energies of reactive intermediates during the water splitting reactions, and for creating more active sites via generating more defects and distortions in the lattice. Heteroatom doping can be divided into two types, namely substitutional doping and surface charge transfer doping, depending on the interactions between heteroatoms and layered TMS.

Non-layered TMS are the most extensively studied bifunctional electrocatalysts for overall water splitting, with the advantages of being widely available, inexpensive, chemically stable, and highly active, and examples include FeS2, Ni3S2, NiS2, CoS2, and Co3S4. Many strategies have been researched to improve their bifunctional electrocatalytic performance via increasing the number of active sites and enhancing the intrinsic activity and electrical conductivity, e.g., composite designing, heteroatom doping, strain engineering, facet engineering, edge engineering, and defect engineering. Similar to layered TMS, the composite designing of non-layered TMS can enhance the charge transfer kinetics and create new active sites due to synergetic effects between the components and the triggering of additional interfaces and defects. Heteroatom doping also can be used to regulate electronic structures to acquire faster electron transfer and optimize the binding energies of intermediates with non-layered TMS. The development of ternary and quaternary TMS-based materials with excellent redox reversibility and high electronic conductivity has created effective substitutes for noble-metal-based electrocatalysts. Facet engineering is strongly dependent on the synthetic methods used for the controllable exposure of the highly active facets of non-layered TMS, and it is favorable for improving the intrinsic activity. Strain engineering and edge engineering can change the electronic character and induce the creation of more active sites, respectively, and these methods are worth further exploration. Finally, as a universal method for optimizing the electrocatalytic performance of non-layered TMS, defect engineering is not a naturally independent method, and it is usually combined with the formation of heterogeneous interfaces, unique architecture design, and heteroatom doping. In principle, most strategies for enhancing the performance of non-layered TMS are also suitable for use with layered TMS whose OER activities need to be further exploited.

Although a lot of progress has been made related to the development of TMS-based bifunctional electrocatalysts, some challenges still exist, as follows:

(1) The electrocatalytic reaction mechanisms related to these bifunctional electrocatalysts in overall water splitting have still not been fully revealed, and more theoretical study efforts should be devoted to uncovering them and providing strong support for exploiting advanced electrocatalysts. On one hand, the detailed reaction mechanism of the OER requires more investigation to uncover its true intrinsic nature. As a sluggish four electron–proton coupled reaction, the several OH*, O*, and OOH* intermediates formed during the OER make it a great challenge to evolve a consolidated reaction mechanism, and different reaction mechanisms for the OER have been proposed by different research groups.24 On the other hand, the HER reaction mechanism at the atomic level is still ambiguous at present, and it requires more in-depth investigation. As is known, the HER mechanism in alkaline electrolytes is totally different from that in acidic media, where it only involves hydrogen binding energetics. The HER mechanism in alkaline media is more complex, including the adsorption/dissociation of water molecules, the adsorption/desorption of hydrogen, and even the OH poisoning of the electrocatalyst surface. To further explore the reaction mechanisms, advanced in situ characterization techniques, such as in situ X-ray absorption spectroscopy (XAS), in situ Raman spectroscopy, and surface interrogation scanning electrochemical microscopy (SI-SECM), should be widely developed and applied to explore dynamic changes in electronic structures, the dynamic evolution of intermediates, and the binding strengths of adsorbates on the surfaces of electrocatalysts. Nowadays, theoretical calculations based on the first-principles calculations provide powerful support when exploring reaction mechanisms at an atomic level. More importantly, multi-scale modelling approaches involving atomic-level models (such as DFT) and energy minimization techniques (molecular dynamics and Monte Carlo models) could be promising for understanding electrocatalytic processes. Combining theoretical calculations with advanced experimental characterization techniques could provide more in-depth understanding of HER and OER mechanisms.

(2) The development of new or combined approaches for further improving the activity of TMS-based electrocatalysts is essential. The direction of efforts to improve the electrocatalytic activities of TMS-based materials is aimed at increasing the number of active sites, improving the electrical conductivity, and optimizing the absorption energies of reaction intermediates. In this review, several strategies for enhancing the HER and OER performances of TMS-based electrocatalysts have been summarized. The approaches of phase engineering, heteroatom doing, strain engineering, and composite designing are favorable for improving the electron transfer characteristics of materials, and the approaches of edge engineering, facet engineering, and defect engineering are beneficial for increasing the number of exposed active sites. The effects of strain engineering, especially on the OER activities of TMS-based materials, need to be further investigated from a mechanistic perspective and examined through experiments. Utilizing multiple approaches simultaneously to realize a maximized enhancement effect is feasible and effective, for example integrating heteroatom doping and strain engineering together41 or integrating heteroatom doping and composite design together.158 In this regard, the synergistic effects of different optimization methods provide a fascinating way to enhance the electrocatalytic activities of TMS-based materials. Examples include preparing highly active heteroatom-doped TMS-based composites with low-dimensional nanostructures via coupling composite designing, edge engineering, and heteroatom doping, and designing metallic-phase TMS-based composites with special morphologies via coupling composite designing, phase engineering, and edge engineering.

(3) Although enhancements in the activities of TMS-based electrocatalysts are fruitful and activities can even be comparable with noble-metal-based electrocatalysts, the stabilities of most bifunctional electrocatalysts cannot meet the needs of industrial applications completely, as they suffer from susceptibility to corrosion and the destruction of structures during long-term water electrolysis. Moreover, the large-scale preparation of TMS-based electrodes is also a bottleneck for industrial applications. Hence, robust materials and electrodes that can be manufactured on a large scale need to be further investigated urgently. On one hand, the creation of TMS-based materials decorated with protective or sacrificial materials could be a simple method to improve corrosion and oxidation resistance. On the other hand, the in situ growth of active TMS on conductive substrates that could be employed as self-supported electrodes is quite a promising approach for dramatically enhancing durability. Strong interactions between active TMS and substrates could dramatically enhance the electrocatalytic durability and electron transfer during long-term service. In situ growth can also avoid the usage of carbon additives and polymer binders, providing more exposed active sites and reducing cost. Broadly speaking, when employing conductive substrates during the preparation process, the ‘bottom-up’ approaches of hydrothermal, solvothermal, chemical vapor deposition, and solid-phase synthesis all exhibit potential for synthesizing self-supported electrodes on a large scale. In this case, the development of reaction devices capable of the large-scale preparation of TMS-based materials is critical for satisfying industrial application demands. Additionally, advanced preparation methods suitable for the large-scale fabrication of these novel electrocatalysts require more attention and development, such as atomic layer deposition, spray pyrolysis, and pulsed laser deposition.

(4) Currently, reported bifunctional electrocatalysts are mostly employed in alkaline electrolytes, while only a few have been applied in acidic and/or neutral media, as the instability of TMS-based electrocatalysts seriously limits their applications.159 However, universal-pH bifunctional electrocatalysts with superior activity and durability in near-neutral media are still required, as these would have great significance for the industrial large-scale development of seawater electrolysis. For example, other binary, ternary, or multiple sulfides with precisely defined shapes, sizes, and compositions, including Sb, Sn, Mn, Re, Cu, and Zn metal based examples, deserve more attention. In addition, bifunctional electrocatalysts for overall urea–water electrolysis are highly desired,160 as they can be applied to efficient H2 generation and the purification of urea-rich wastewater to meet the requirements of both energy conversion and environmental protection.

All in all, developing bifunctional TMS-based electrocatalysts for overall water splitting is very appealing yet challenging. With continuing research efforts being focused on the strategies mentioned above, it is conceivable that TMS-based bifunctional electrocatalysts with extremely improved activity and long-term stability will be developed for industrial applications.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This work was supported by the Basic Science Center Project of the NSFC (51788104) and National Key Research and Development Program of China (2017YFB1104300).

References

  1. Z. W. Seh, J. Kibsgaard, C. F. Dickens, I. Chorkendorff, J. K. Norskov and T. F. Jaramillo, Combining theory and experiment in electrocatalysis: insights into materials design, Science, 2017, 355, 4998 CrossRef .
  2. S. Chandrasekaran, L. Yao, L. Deng, C. Bowen, Y. Zhang, S. Chen, Z. Lin, F. Peng and P. Zhang, Recent advances in metal sulfides: from controlled fabrication to electrocatalytic, photocatalytic and photoelectrochemical water splitting and beyond, Chem. Soc. Rev., 2019, 48, 4178–4280 RSC .
  3. J. W. Joo, T. Kim, J. Lee, S. I. Choi and K. Lee, Morphology-controlled metal sulfides and phosphides for electrochemical water splitting, Adv. Mater., 2019, 37, 1806682 CrossRef .
  4. J. S. Jirkovsky, C. D. Malliakas, P. P. Lopes, N. Danilovic, S. S. Kota, K. C. Chang, B. Genorio, D. Strmcnik, V. R. Stamenkovic, M. G. Kanatzidis and N. M. Markovic, Design of active and stable Co-Mo-Sx chalcogels as pH-universal catalysts for the hydrogen evolution reaction, Nat. Mater., 2016, 15, 197–203 CrossRef .
  5. H. Jin, J. Joo, N. K. Chaudhari, S. I. Choi and K. Lee, Recent progress in bifunctional electrocatalysts for overall water splitting under acidic conditions, ChemElectroChem, 2019, 6, 3244–3253 CrossRef CAS .
  6. B. Tang, X. Yang, Z. Kang and L. Feng, Crystallized RuTe2 as unexpected bifunctional catalyst for overall water splitting, Appl. Catal. B Environ., 2020, 278, 119281 CrossRef CAS .
  7. Y. Li, Y. Sun, Y. Qin, W. Zhang, L. Wang, M. Luo, H. Yang and S. Guo, Recent advances on water-splitting electrocatalysis mediated by noble-metal-based nanostructured materials, Adv. Energy Mater., 2020, 10, 1903120 CrossRef CAS .
  8. G. Zhang, Y. S. Feng, W. T. Lu, D. He, C. Y. Wang, Y. K. Li, X. Y. Wang and F. F. Cao, Enhanced catalysis of electrochemical overall water splitting in alkaline media by Fe doping in Ni3S2 nanosheet arrays, ACS Catal., 2018, 8, 5431–5441 CrossRef CAS .
  9. J. R. Varcoe, P. Atanassov, D. R. Dekel, A. M. Herring, M. A. Hickner, P. A. Kohl, A. R. Kucernak, W. E. Mustain, K. Nijmeijer, K. Scott, T. Xu and L. Zhuang, Anion-exchange membranes in electrochemical energy systems, Energy Environ. Sci., 2014, 7, 3135–3191 RSC .
  10. Z. Kang, H. Guo, J. Wu, X. Sun, Z. Zhang, Q. Liao, S. Zhang, H. Si, P. Wu, L. Wang and Y. Zhang, Engineering an earth-abundant element-based bifunctional electrocatalyst for highly efficient and durable overall water splitting, Adv. Funct. Mater., 2019, 29, 1807031 CrossRef .
  11. J. Wang, X. Yue, Y. Yang, S. Sirisomboonchai, P. Wang, X. Ma, A. Abudula and G. Guan, Earth-abundant transition-metal-based bifunctional catalysts for overall electrochemical water splitting: a review, J. Alloys Compd., 2020, 819, 153346 CrossRef CAS .
  12. F. Yu, L. Yu, I. K. Mishra, Y. Yu, Z. F. Ren and H. Q. Zhou, Recent developments in earth-abundant and non-noble electrocatalysts for water electrolysis, Mater. Today Phys., 2018, 7, 121–138 CrossRef .
  13. C. Hu, L. Zhang and J. Gong, Recent progress made in the mechanism comprehension and design of electrocatalysts for alkaline water splitting, Energy Environ. Sci., 2019, 12, 2620–2645 RSC .
  14. S. Anantharaj, S. R. Ede, K. Sakthikumar, K. Karthick, S. Mishra and S. Kundu, Recent trends and perspectives in electrochemical water splitting with an emphasis on sulfide, selenide, and phosphide catalysts of Fe, Co, and Ni: a review, ACS Catal., 2016, 6, 8069–8097 CrossRef CAS .
  15. Q. Lu, Y. Yu, Q. Ma, B. Chen and H. Zhang, 2D Transition-Metal-Dichalcogenide-Nanosheet-Based Composites for Photocatalytic and Electrocatalytic Hydrogen Evolution Reactions, Adv. Mater., 2016, 28, 1917–1933 CrossRef CAS .
  16. X. Chia, A. Ambrosi, P. Lazar, Z. Sofer and M. Pumera, Electrocatalysis of layered Group 5 metallic transition metal dichalcogenides (MX2, M = V, Nb, and Ta; X = S, Se, and Te), J. Mater. Chem. A, 2016, 4, 14241–14253 RSC .
  17. C. Feng, M. B. Faheem, J. Fu, Y. Xiao, C. Li and Y. Li, Fe-Based Electrocatalysts for Oxygen Evolution Reaction: Progress and Perspectives, ACS Catal., 2020, 10, 4019–4047 CrossRef CAS .
  18. W. Zhang, L. Cui and J. Liu, Recent advances in cobalt-based electrocatalysts for hydrogen and oxygen evolution reactions, J. Alloys Compd., 2020, 821, 153542 CrossRef CAS .
  19. Y. N. Guo, T. Park, J. W. Yi, J. Henzie, J. Kim, Z. L. Wang, B. Jiang, Y. Bando, Y. Sugahara, J. Tang and Y. Yamauchi, Nanoarchitectonics for transition-metal-sulfide-based electrocatalysts for water splitting, Adv. Mater., 2019, 31, 1807134 CrossRef .
  20. X. Zou and Y. Zhang, Noble metal-free hydrogen evolution catalysts for water splitting, Chem. Soc. Rev., 2015, 44, 5148–5180 RSC .
  21. H. Sun, Z. Yan, F. Liu, W. Xu, F. Cheng and J. Chen, Self-supported transition-metal-based electrocatalysts for hydrogen and oxygen evolution, Adv. Mater., 2019, 1806326 Search PubMed .
  22. N. Mahmood, Y. Yao, J. W. Zhang, L. Pan, X. Zhang and J. J. Zou, Electrocatalysts for hydrogen evolution in alkaline electrolytes: mechanisms, challenges, and prospective solutions, Adv. Sci., 2018, 5, 1700464 CrossRef .
  23. K. Ojha, S. Saha, P. Dagar and A. K. Ganguli, Nanocatalysts for hydrogen evolution reactions, Phys. Chem. Chem. Phys., 2018, 20, 6777–6799 RSC .
  24. N. T. Suen, S. F. Hung, Q. Quan, N. Zhang, Y. J. Xu and H. M. Chen, Electrocatalysis for the oxygen evolution reaction: recent development and future perspectives, Chem. Soc. Rev., 2017, 46, 337–365 RSC .
  25. Z. Liu, C. Zhang, H. Liu and L. Feng, Efficient synergism of NiSe2 nanoparticle/NiO nanosheet for energy-relevant water and urea electrocatalysis, Appl. Catal. B Environ., 2020, 276, 119281 CrossRef .
  26. J. Rossmeisl, Z. W. Qu, H. Zhu, G. J. Kroes and J. K. Nørskov, Electrolysis of water on oxide surfaces, J. Electroanal. Chem., 2007, 607, 83–89 CrossRef CAS .
  27. M. Zeng and Y. Li, Recent advances in heterogeneous electrocatalysts for the hydrogen evolution reaction, J. Mater. Chem. A, 2015, 3, 14942–14962 RSC .
  28. C. C. McCrory, S. Jung, J. C. Peters and T. F. Jaramillo, Benchmarking heterogeneous electrocatalysts for the oxygen evolution reaction, J. Am. Chem. Soc., 2013, 135, 16977–16987 CrossRef CAS .
  29. D. Li, H. Liu and L. Feng, A Review on Advanced FeNi-Based Catalysts for Water Splitting Reaction, Energy Fuels, 2020, 34, 13491–13522 CrossRef CAS .
  30. C. C. McCrory, S. Jung, J. C. Peters and T. F. Jaramillo, Benchmarking heterogeneous electrocatalysts for the oxygen evolution reaction, J. Am. Chem. Soc., 2013, 135, 16977–16987 CrossRef CAS .
  31. M. Wang, L. Zhang, M. Huang, Y. Liu, Y. Zhong, J. Pan, Y. Wang and H. Zhu, Morphology-controlled tantalum diselenide structures as self-optimizing hydrogen evolution catalysts, Energy. Environ. Mater., 2019, 3, 12–18 CrossRef .
  32. M. Chhowalla, H. S. Shin, G. Eda, L. J. Li, K. P. Loh and H. Zhang, The chemistry of two-dimensional layered transition metal dichalcogenide nanosheets, Nat. Chem., 2013, 5, 263–275 CrossRef .
  33. A. Ambrosi and M. Pumera, Exfoliation of layered materials using electrochemistry, Chem. Soc. Rev., 2018, 47, 7213–7224 RSC .
  34. X. Chia, A. Y. Eng, A. Ambrosi, S. M. Tan and M. Pumera, Electrochemistry of nanostructured layered transition-metal dichalcogenides, Chem. Rev., 2015, 115, 11941–11966 CrossRef CAS .
  35. G. H. Han, D. L. Duong, D. H. Keum, S. J. Yun and Y. H. Lee, van der Waals metallic transition metal dichalcogenides, Chem. Rev., 2018, 118, 6297–6336 CrossRef CAS .
  36. M. A. Lukowski, A. S. Daniel, F. Meng, A. Forticaux, L. Li and S. Jin, Enhanced hydrogen evolution catalysis from chemically exfoliated metallic MoS2 nanosheets, J. Am. Chem. Soc., 2013, 135, 10274–10277 CrossRef CAS .
  37. A. Ambrosi, Z. Sofer and M. Pumera, 2H-1T phase transition and hydrogen evolution activity of MoS2, MoSe2, WS2 and WSe2 strongly depends on the MX2 composition, Chem. Commun., 2015, 51, 8450–8453 RSC .
  38. D. Kong, J. J. Cha, H. Wang, H. R. Lee and Y. Cui, First-row transition metal dichalcogenide catalysts for hydrogen evolution reaction, Energy Environ. Sci., 2013, 6, 3553–3558 RSC .
  39. C. R. Zhu, D. Gao, J. Ding, D. Chao and J. Wang, TMD-based highly efficient electrocatalysts developed by combined computational and experimental approaches, Chem. Soc. Rev., 2018, 47, 4332–4356 RSC .
  40. T. H. M. Lau, S. Wu, R. Kato, T. S. Wu, J. Kulhavý, J. Mo, J. Zheng, J. S. Foord, Y. L. Soo, K. Suenaga, M. T. Darby and S. C. E. Tsang, Engineering monolayer 1T-MoS2 into a bifunctional electrocatalyst via sonochemical doping of isolated transition metal atoms, ACS Catal., 2019, 9, 7527–7534 CrossRef CAS .
  41. H. Li, C. Tsai, A. L. Koh, L. Cai, A. W. Contryman, A. H. Fragapane, J. Zhao, H. S. Han, H. C. Manoharan, F. A. Pedersen, J. K. Norskov and X. Zheng, Activating and optimizing MoS2 basal planes for hydrogen evolution through the formation of strained sulphur vacancies, Nat. Mater., 2016, 15, 364 CrossRef CAS .
  42. J. Xie, J. Zhang, S. Li, F. Grote, X. Zhang, H. Zhang, R. Wang, Y. Lei, B. Pan and Y. Xie, Controllable disorder engineering in oxygen-incorporated MoS2 ultrathin nanosheets for efficient hydrogen evolution, J. Am. Chem. Soc., 2013, 135, 17881–17888 CrossRef CAS .
  43. K. Sun, L. Zeng, S. Liu, L. Zhao, H. Zhu, J. Zhao, Z. Liu, D. Cao, Y. Hou, Y. Liu, Y. Pan and C. Liu, Design of basal plane active MoS2 through one-step nitrogen and phosphorus co-doping as an efficient pH-universal electrocatalyst for hydrogen evolution, Nano Energy, 2019, 58, 862–869 CrossRef CAS .
  44. X. Chen, Z. Wang, Y. Qiu, J. Zhang, G. Liu, W. Zheng, W. Feng, W. Cao, P. Hu and W. Hu, Controlled growth of vertical 3D MoS2(1−x)Se2x nanosheets for an efficient and stable hydrogen evolution reaction, J. Mater. Chem. A, 2016, 4, 18060–18066 RSC .
  45. P. Liu, J. Zhu, J. Zhang, K. Tao, D. Gao and P. Xi, Active basal plane catalytic activity and conductivity in Zn doped MoS2 nanosheets for efficient hydrogen evolution, Electrochim. Acta, 2018, 260, 24–30 CrossRef CAS .
  46. Y. Liu, J. Wu, K. P. Hackenberg, J. Zhang, Y. M. Wang, Y. Yang, K. Keyshar, J. Gu, T. Ogitsu, R. Vajtai, J. Lou, P. M. Ajayan, B. C. Wood and B. I. Yakobson, Self-optimizing, highly surface-active layered metal dichalcogenide catalysts for hydrogen evolution, Nat. Energy, 2017, 2, 17127 CrossRef CAS .
  47. R. Sun, M. K. Y. Chan and G. Ceder, First-principles electronic structure and relative stability of pyrite and marcasite: Implications for photovoltaic performance, Phys. Rev. B: Condens. Matter Mater. Phys., 2011, 83, 235311 CrossRef .
  48. M. R. Gao, Y. R. Zheng, J. Jiang and S. H. Yu, Pyrite-type nanomaterials for advanced electrocatalysis, Acc. Chem. Res., 2017, 50, 2194–2204 CrossRef CAS .
  49. A. N. Buckley and R. Woods, Electrochemical and XPS studies of the surface oxidation of synthetic heazlewoodite (Ni3S2), J. Appl. Electrochem., 1991, 21, 575–582 CrossRef CAS .
  50. J. Dong, F. Q. Zhang, Y. Yang, Y. B. Zhang, H. He, X. Huang, X. Fan and X. M. Zhang, (003)-Facet-exposed Ni3S2 nanoporous thin films on nickel foil for efficient water splitting, Appl. Catal. B Environ., 2019, 243, 693–702 CrossRef CAS .
  51. Y. Chen, K. Yang, B. Jiang, J. Li, M. Zeng and L. Fu, Emerging two-dimensional nanomaterials for electrochemical hydrogen evolution, J. Mater. Chem. A, 2017, 5, 8187–8208 RSC .
  52. T. F. Jaramillo, K. P. Jørgensen, J. Bonde, J. H. Nielsen, S. Horch and I. Chorkendorff, Identification of Active Edge Sites for Electrochemical H2 Evolution from MoS2 Nanocatalysts, Science, 2007, 317, 100–102 CrossRef CAS .
  53. R. Subbaraman, D. Tripkovic, K. C. Chang, D. Strmcnik, A. P. Paulikas, P. Hirunsit, M. Chan, J. Greeley, V. Stamenkovic and N. M. Markovic, Trends in activity for the water electrolyser reactions on 3d M (Ni, Co, Fe, Mn) hydr(oxy)oxide catalysts, Nat. Mater., 2012, 11, 550–557 CrossRef CAS .
  54. S. Jin, Are Metal Chalcogenides, Nitrides, and Phosphides Oxygen Evolution Catalysts or Bifunctional Catalysts?, ACS Energy Lett., 2017, 2, 1937–1938 CrossRef CAS .
  55. O. Mabayoje, A. Shoola, B. R. Wygant and C. B. Mullins, The Role of Anions in Metal Chalcogenide Oxygen Evolution Catalysis: Electrodeposited Thin Films of Nickel Sulfide as “Pre-catalysts”, ACS Energy Lett., 2016, 1, 195–201 CrossRef CAS .
  56. S. A. Han, R. Bhatia and S. W. Kim, Synthesis, properties and potential applications of two-dimensional transition metal dichalcogenides, Nano Converg., 2015, 2, 17 CrossRef .
  57. H. Li, J. Wu, Z. Yin and H. Zhang, Preparation and applications of mechanically exfoliated single-layer and multilayer MoS2 and WSe2 nanosheets, Acc. Chem. Res., 2014, 47, 1067–1075 CrossRef CAS .
  58. Z. Li, R. J. Young, C. Backes, W. Zhao, X. Zhang, A. A. Zhukov, E. Tillotson, A. P. Conlan, F. Ding, S. J. Haigh, K. S. Novoselov and J. N. Coleman, Mechanisms of Liquid-Phase Exfoliation for the Production of Graphene, ACS Nano, 2020, 14, 10976–10985 CrossRef CAS .
  59. J. N. Coleman, M. Lotya, A. O'Neill, S. D. Bergin, P. J. King, U. Khan and K. Young, Two-dimensional nanosheets produced by liquid exfoliation of layered materials, Science, 2011, 331, 568–571 CrossRef CAS .
  60. E. Varrla, C. Backes, K. R. Paton, A. Harvey, Z. Gholamvand, J. McCauley and J. N. Coleman, Large-scale production of size-controlled MoS2 nanosheets by shear exfoliation, Chem. Mater., 2015, 27, 1129–1139 CrossRef CAS .
  61. X. Zhao, X. Ma, J. Sun, D. Li and X. Yang, Enhanced catalytic activities of surfactant-assisted exfoliated WS2 nanodots for hydrogen evolution, ACS Nano, 2016, 10, 2159–2166 CrossRef CAS .
  62. X. Fan, P. Xu, Y. C. Li, D. Zhou, Y. Sun, M. A. Nguyen, M. Terrones and T. E. Mallouk, Controlled exfoliation of MoS2 crystals into trilayer nanosheets, J. Am. Chem. Soc., 2016, 138, 5143–5149 CrossRef CAS .
  63. N. Liu, P. Kim, J. H. Kim, J. H. Ye, S. Kim and C. J. Lee, Large-area atomically thin MoS2 nanosheets prepared using electrochemical exfoliation, ACS Nano, 2014, 8, 6902–6910 CrossRef CAS .
  64. X. Huang, Z. Zeng and H. Zhang, Metal dichalcogenide nanosheets: preparation, properties and applications, Chem. Soc. Rev., 2013, 42, 1934–1946 RSC .
  65. S. C. Han, H. S. Kim, M. S. Song, J. H. Kim, H. J. Ahn and J. Y. Lee, Nickel sulfide synthesized by ball milling as an attractive cathode material for rechargeable lithium batteries, J. Alloys Compd., 2003, 351, 273–278 CrossRef CAS .
  66. A. Ambrosi, X. Chia, Z. Sofer and M. Pumera, Enhancement of electrochemical and catalytic properties of MoS2 through ball-milling, Electrochem. Commun., 2015, 54, 36–40 CrossRef CAS .
  67. P. Luo, H. Zhang, L. Liu, Y. Zhang, J. Deng, C. Xu, N. Hu and Y. Wang, Targeted synthesis of unique nickel sulfide (NiS, NiS2) microarchitectures and the applications for the enhanced water splitting system, ACS Appl. Mater. Interfaces, 2017, 9, 2500–2508 CrossRef CAS .
  68. W. Dong, X. Wang, B. Li, L. Wang, B. Chen, C. Li, X. Li, T. Zhang and Z. Shi, Hydrothermal synthesis and structure evolution of hierarchical cobalt sulfide nanostructures, Dalton Trans., 2011, 40, 243–248 RSC .
  69. Q. Xiong, X. Zhang, H. Wang, G. Liu, G. Wang, H. Zhang and H. Zhao, One-step synthesis of cobalt-doped MoS2 nanosheets as bifunctional electrocatalysts for overall water splitting under both acidic and alkaline conditions, Chem. Commun., 2018, 54, 3859–3862 RSC .
  70. Y. Li, H. Wang, L. Xie, Y. Liang, G. Hong and H. Dai, MoS2 nanoparticles grown on graphene: an advanced catalyst for the hydrogen evolution reaction, J. Am. Chem. Soc., 2011, 133, 7296–7299 CrossRef CAS .
  71. C. Z. Yuan, Z. T. Sun, Y. F. Jiang, Z. K. Yang, N. Jiang, Z. W. Zhao, U. Y. Qazi, W. H. Zhang and A. W. Xu, 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, 1604161 CrossRef .
  72. Q. Wang, L. Jiao, H. Du, W. Peng, Y. Han, D. Song, Y. Si, Y. Wang and H. Yuan, Novel flower-like CoS hierarchitectures: one-pot synthesis and electrochemical properties, J. Mater. Chem., 2011, 21, 327–329 RSC .
  73. J. Yuan, J. Wu, W. J. Hardy, P. Loya, M. Lou, Y. Yang, S. Najmaei, M. Jiang, F. Qin, K. Keyshar, H. Ji, W. Gao, J. Bao, J. Kono, D. Natelson, P. M. Ajayan and J. Lou, Facile synthesis of single cystal vanadium disulfide nanosheets by chemical vapor deposition for efficient hydrogen evolution reaction, Adv. Mater., 2015, 27, 5605–5609 CrossRef CAS .
  74. L. Samad, M. C. Acevedo, M. J. Shearer, K. Park, R. J. Hamers and S. Jin, Direct chemical vapor deposition synthesis of phase-pure iron pyrite (FeS2) thin films, Chem. Mater., 2015, 27, 3108–3114 CrossRef CAS .
  75. J. Shi, X. Wang, S. Zhang, L. Xiao, Y. Huan, Y. Gong, Z. Zhang, Y. Li, X. Zhou, M. Hong, Q. Fang, Q. Zhang, X. Liu, L. Gu, Z. Liu and Y. Zhang, Two-dimensional metallic tantalum disulfide as a hydrogen evolution catalyst, Nat. Commun., 2017, 8, 958 CrossRef .
  76. H. U. Kim, V. Kanade, M. Kim, K. S. Kim, B. S. An, H. Seok, H. Yoo, L. E. Chaney, S. I. Kim, C. W. Yang, G. Y. Yeom, D. Whang, J. H. Lee and T. Kim, Wafer-scale and low-temperature growth of 1T-WS2 film for efficient and stable hydrogen evolution reaction, Small, 2020, 16, 1905000 CrossRef CAS .
  77. I. Song, C. Park, M. Hong, J. Baik, H. J. Shin and H. C. Choi, Patternable large-scale molybdenium disulfide atomic layers grown by gold-assisted chemical vapor deposition, Angew. Chem., Int. Ed., 2014, 126, 1290–1293 CrossRef .
  78. M. Wang, L. Zhang, M. Huang, Q. Zhang, X. Zhao, Y. He, S. Lin, J. Pan and H. Zhu, One-step synthesis of a hierarchical self-supported WS2 film for efficient electrocatalytic hydrogen evolution, J. Mater. Chem. A, 2019, 7, 22405–22411 RSC .
  79. M. C. Acevedo, M. S. Faber, Y. Tan, R. J. Hamers and S. Jin, Synthesis and properties of semiconducting iron pyrite (FeS2) nanowires, Nano Lett., 2012, 12, 1977–1982 CrossRef .
  80. M. S. Faber, M. A. Lukowski, Q. Ding, N. S. Kaiser and S. Jin, Earth-abundant metal pyrites (FeS2, CoS2, NiS2, and their alloys) for highly efficient hydrogen evolution and polysulfide reduction electrocatalysis, J. Phys. Chem. C, 2014, 118, 21347–21356 CrossRef CAS .
  81. R. Miao, B. Dutta, S. Sahoo, J. He, W. Zhong, S. A. Cetegen, T. Jiang, S. P. Alpay and S. L. Suib, Mesoporous Iron Sulfide for Highly Efficient Electrocatalytic Hydrogen Evolution, J. Am. Chem. Soc., 2017, 139, 13604–13607 CrossRef CAS .
  82. M. C. Acevedo, D. Liang, K. S. Chew, J. P. DeGrave, N. S. Kaiser and S. Jin, Synthesis, Characterization, and Variable Range Hopping Transport of Pyrite (FeS2) Nanorods, Nanobelts, and Nanoplates, ACS Nano, 2013, 7(2), 1731–1739 CrossRef .
  83. C. Bara, A. F. L. Humblot, E. Fonda, A. S. Gay, A. L. Taleb, E. Devers, M. Digne, G. D. Pirngruber and X. Carrier, Surface-dependent sulfidation and orientation of MoS2 slabs on alumina-supported model hydrodesulfurization catalysts, J. Catal., 2016, 344, 591–605 CrossRef CAS .
  84. X. M. Geng, W. W. Sun, W. Wu, B. Chen, A. A. Hilo, M. Benamara, H. L. Zhu, F. Watanabe, J. B. Cui and T. Chen, Pure and stable metallic phase molybdenum disulfide nanosheets for hydrogen evolution reaction, Nat. Commun., 2016, 7, 10672 CrossRef CAS .
  85. H. Li, Y. Tan, P. Liu, C. Guo, M. Luo, J. Han, T. Lin, F. Huang and M. Chen, Atomic-Sized Pores Enhanced Electrocatalysis of TaS2 Nanosheets for Hydrogen Evolution, Adv. Mater., 2016, 28, 8945–8949 CrossRef CAS .
  86. L. Najafi, S. Bellani, R. O. Nuñez, B. M. García, M. Prato, V. Mazánek, D. Debellis, S. Lauciello, R. Brescia, Z. Sofer and F. Bonaccorso, Niobium disulphide (NbS2)-based heterogeneous electrocatalysts for an efficient hydrogen evolution reaction, J. Mater. Chem. A, 2019, 7, 25593–25608 RSC .
  87. J. Lin, Z. Peng, G. Wang, D. Zakhidov, E. Larios, M. J. Yacaman and J. M. Tour, Enhanced Electrocatalysis for Hydrogen Evolution Reactions from WS2 Nanoribbons, Adv. Energy Mater., 2014, 4, 1301875 CrossRef .
  88. Q. Tang and D. Jiang, Mechanism of Hydrogen Evolution Reaction on 1T-MoS2 from First Principles, ACS Catal., 2016, 6, 4953–4961 CrossRef CAS .
  89. M. Zhang, Y. He, D. Yan, H. Xu, A. Wang, Z. Chen, S. Wang, H. Luo and K. Yan, Multifunctional 2H-TaS2 nanoflakes for efficient supercapacitors and electrocatalytic evolution of hydrogen and oxygen, Nanoscale, 2019, 11, 22255–22260 RSC .
  90. K. Yan and Y. Lu, Direct Growth of MoS2 Microspheres on Ni Foam as a Hybrid Nanocomposite Efficient for Oxygen Evolution Reaction, Small, 2016, 12, 2975–2981 CrossRef CAS .
  91. J. Wu, M. Liu, K. Chatterjee, K. P. Hackenberg, J. Shen, X. Zou, Y. Yan, J. Gu, Y. Yang, J. Lou and P. M. Ajayan, Exfoliated 2D Transition Metal Disulfides for Enhanced Electrocatalysis of Oxygen Evolution Reaction in Acidic Medium, Adv. Mater. Interfaces, 2016, 3, 1500669 CrossRef .
  92. S. Wei, X. Cui, Y. Xu, B. Shang, Q. Zhang, L. Gu, X. Fan, L. Zheng, C. Hou, H. Huang, S. Wen and W. Zheng, Iridium-Triggered Phase Transition of MoS2 Nanosheets Boosts Overall Water Splitting in Alkaline Media, ACS Energy Lett., 2019, 4, 368–374 CrossRef CAS .
  93. J. Zhang, T. Wang, D. Pohl, B. Rellinghaus, R. Dong, S. Liu, X. Zhuang and X. Feng, Interface Engineering of MoS2/Ni3S2 Heterostructures for Highly Enhanced Electrochemical Overall-Water-Splitting Activity, Angew. Chem., Int. Ed., 2016, 55, 6702–6707 CrossRef CAS .
  94. S. Peng, L. Li, J. Zhang, T. L. Tan, T. Zhang, D. Ji, X. Han, F. Cheng and S. Ramakrishna, Engineering Co9S8/WS2 array films as bifunctional electrocatalysts for efficient water splitting, J. Mater. Chem. A, 2017, 5, 23361–23368 RSC .
  95. J. Hou, B. Zhang, Z. Li, S. Cao, Y. Sun, Y. Wu, Z. Gao and L. Sun, Vertically Aligned Oxygenated-CoS2-MoS2 Heteronanosheet Architecture from Polyoxometalate for Efficient and Stable Overall Water Splitting, ACS Catal., 2018, 8, 4612–4621 CrossRef CAS .
  96. J. Liu, J. Wang, B. Zhang, Y. Ruan, H. Wan, X. Ji, K. Xu, D. Zha, L. Miao and J. Jiang, Mutually beneficial Co3O4@MoS2 heterostructures as a highly efficient bifunctional catalyst for electrochemical overall water splittin, J. Mater. Chem. A, 2018, 6, 2067–2072 RSC .
  97. Y. Yang, H. Yao, Z. Yu, S. M. Islam, H. He, M. Yuan, Y. Yue, K. Xu, W. Hao, G. Sun, H. Li, S. Ma, P. Zapol and M. G. Kanatzidis, Hierarchical Nanoassembly of MoS2/Co9S8/Ni3S2/Ni as a Highly Efficient Electrocatalyst for Overall Water Splitting in a Wide pH Range, J. Am. Chem. Soc., 2019, 141, 10417–10430 CrossRef CAS .
  98. D. Wang, Q. Li, C. Han, Z. Xing and X. Yang, When NiO@Ni Meets WS2 Nanosheet Array: A Highly Efficient and Ultrastable Electrocatalyst for Overall Water Splitting, ACS Cent. Sci., 2018, 4, 112–119 CrossRef CAS .
  99. D. Ji, S. Peng, L. Fan, L. Li, X. Qin and S. Ramakrishna, Thin MoS2 nanosheets grafted MOFs-derived porous Co-N-C flakes grown on electrospun carbon nanofibers as self-supported bifunctional catalysts for overall water splitting, J. Mater. Chem. A, 2017, 5, 23898–23908 RSC .
  100. B. Shang, P. Ma, J. Fan, L. Jiao, Z. Liu, Z. Zhang, N. Chen, Z. Cheng, X. Cui and W. Zheng, Stabilized monolayer 1T MoS2 embedded in CoOOH for highly efficient overall water splitting, Nanoscale, 2018, 10, 12330–12336 RSC .
  101. C. Xie, D. Yan, W. Chen, Y. Zou, R. Chen, S. Zang, Y. Wang, X. Yao and S. Wang, Insight into the design of defect electrocatalysts: From electronic structure to adsorption energy, Mater. Today, 2019, 31, 47–68 CrossRef CAS .
  102. J. Deng, H. Li, J. Xiao, Y. Tu, D. Deng, H. Yang, H. Tian, J. Li, P. Ren and X. Bao, Triggering the electrocatalytic hydrogen evolution activity of the inert two-dimensional MoS2 surface via single-atom metal doping, Energy Environ. Sci., 2015, 8, 1594–1601 RSC .
  103. V. P. Pham and G. Y. Yeom, Recent Advances in Doping of Molybdenum Disulfide: Industrial Applications and Future Prospects, Adv. Mater., 2016, 28, 9024–9059 CrossRef CAS .
  104. Q. Xiong, Y. Wang, P. F. Liu, L. R. Zheng, G. Wang, H. G. Yang, P. K. Wong, H. Zhang and H. Zhao, Cobalt Covalent Doping in MoS2 to Induce Bifunctionality of Overall Water Splitting, Adv. Mater., 2018, 30, 1801450 CrossRef .
  105. U. N. Pan, T. I. Singh, D. R. Paudel, C. C. Gudal, N. H. Kim and J. H. Lee, Covalent doping of Ni and P on 1T-enriched MoS2 bifunctional 2D-nanostructures with active basal planes and expanded interlayers boosts electrocatalytic water splitting, J. Mater. Chem. A, 2020, 8, 19654–19664 RSC .
  106. J. Y. Xue, F. L. Li, Z. Y. Zhao, C. Li, C. Y. Ni, H. W. Gu, D. J. Young and J. P. Lang, In Situ Generation of Bifunctional Fe-Doped MoS2 Nanocanopies for Efficient Electrocatalytic Water Splitting, Inorg. Chem., 2019, 58, 11202–11209 CrossRef CAS .
  107. X. Xu, H. Xu and D. Cheng, Design of high-performance MoS2 edge supported single-metal atom bifunctional catalysts for overall water splitting via a simple equation, Nanoscale, 2019, 11, 20228–20237 RSC .
  108. I. S. Kwon, T. T. Debela, I. H. Kwak, Y. C. Park, J. Seo, J. Y. Shim, S. J. Yoo, J. G. Kim, J. Park and H. S. Kang, Ruthenium Nanoparticles on Cobalt-Doped 1T' Phase MoS2 Nanosheets for Overall Water Splitting, Small, 2020, 16, 2000081 CrossRef CAS .
  109. N. Jiang, Q. Tang, M. L. Sheng, B. You, D. Jiang and Y. J. Sun, Nickel sulfides for electrocatalytic hydrogen evolution under alkaline conditions: a case study of crystalline NiS, NiS2, and Ni3S2 nanoparticles, Catal. Sci. Technol., 2016, 6, 1077–1084 RSC .
  110. X. Shi, X. Ling, L. Li, C. Zhong, Y. Deng, X. Han and W. Hu, Nanosheets assembled into nickel sulfide nanospheres with enriched Ni3+ active sites for efficient water-splitting and zinc-air batteries, J. Mater. Chem. A, 2019, 7, 23787–23793 RSC .
  111. X. Zheng, X. Han, Y. Zhang, J. Wang, C. Zhong, Y. Deng and W. Hu, Controllable synthesis of nickel sulfide nanocatalysts and their phase-dependent performance for overall water splitting, Nanoscale, 2019, 11, 5646–5654 RSC .
  112. L. Zeng, K. Sun, Z. Yang, S. Xie, Y. Chen, Z. Liu, Y. Liu, J. Zhao, Y. Liu and C. Liu, Tunable 3D hierarchical Ni3S2 superstructures as efficient and stable bifunctional electrocatalysts for both H2 and O2 generation, J. Mater. Chem. A, 2018, 6, 4485–4493 RSC .
  113. J. Zhang, Y. Li, T. Zhu, Y. Wang, J. Cui, J. Wu, H. Xu, X. Shu, Y. Qin, H. Zheng, P. M. Ajayan, Y. Zhang and Y. Wu, 3D Coral-Like Ni3S2 on Ni Foam as a Bifunctional Electrocatalyst for Overall Water Splitting, ACS Appl. Mater. Interfaces, 2018, 10, 31330–31339 CrossRef CAS .
  114. W. Zhu, X. Yue, W. Zhang, S. Yu, Y. Zhang, J. Wang and J. Wang, Nickel sulfide microsphere film on Ni foam as an efficient bifunctional electrocatalyst for overall water splitting, Chem. Commun., 2016, 52, 1486–1489 RSC .
  115. X. Ma, W. Zhang, Y. Deng, C. Zhong, W. Hu and X. Han, Phase and composition controlled synthesis of cobalt sulfide hollow nanospheres for electrocatalytic water splitting, Nanoscale, 2018, 10, 4816–4824 RSC .
  116. M. Fronzi, M. H. N. Assadi and M. J. Ford, Ab Initio Investigation of Water Adsorption and Hydrogen Evolution on Co9S8 and Co3S4 Low-Index Surfaces, ACS Omega, 2018, 3, 12215–12228 CrossRef CAS .
  117. Z. Li, M. Xiao, Y. Zhou, D. Zhang, H. Wang, X. Liu, D. Wang and W. Wang, Pyrite FeS2/C nanoparticles as an efficient bi-functional catalyst for overall water splitting, Dalton Trans., 2018, 47, 14917–14923 RSC .
  118. H. Li, S. Chen, Y. Zhang, Q. Zhang, X. Jia, Q. Zhang, L. Gu, X. Sun, L. Song and X. Wang, Systematic design of superaerophobic nanotube-array electrode comprised of transition-metal sulfides for overall water splitting, Nat. Commun., 2018, 9, 2452 CrossRef .
  119. Y. Li, J. Yin, L. An, M. Lu, K. Sun, Y. Q. Zhao, D. Gao, F. Cheng and P. Xi, FeS2/CoS2 Interface Nanosheets as Efficient Bifunctional Electrocatalyst for Overall Water Splitting, Small, 2018, 14, 1801070 CrossRef .
  120. H. Zhu, J. Zhang, R. Y. Zhang, M. Du, Q. Wang, G. Gao, J. Wu, G. Wu, M. Zhang, B. Liu, J. Yao and X. Zhang, When cubic cobalt sulfide meets layered molybdenum disulfide: a core-shell system toward synergetic electrocatalytic water splitting, Adv. Mater., 2015, 27, 4752–4759 CrossRef CAS .
  121. J. Lin, H. Wang, X. Zheng, Y. Du, C. Zhao, J. Qi, J. Cao, W. Fei and J. Feng, Controllable synthesis of core-branch Ni3S2/Co9S8 directly on nickel foam as an efficient bifunctional electrocatalyst for overall water splitting, J. Power Sources, 2018, 401, 329–335 CrossRef CAS .
  122. W. Xin, W. J. Jiang, Y. Lian, H. Li, S. Hong, S. Xu, H. Yan and J. S. Hu, NiS2 nanodotted carnation-like CoS2 for enhanced electrocatalytic water splitting, Chem. Commun., 2019, 55, 3781–3784 RSC .
  123. Q. Li, D. Wang, C. Han, X. Ma, Q. Lu, Z. Xing and X. Yang, Construction of amorphous interface in an interwoven NiS/NiS2 structure for enhanced overall water splitting, J. Mater. Chem. A, 2018, 6, 8233–8237 RSC .
  124. F. Jing, Q. Lv, J. Xiao, Q. Wang and S. Wang, Highly active and dual-function self-supported multiphase NiS-NiS2-Ni3S2/NF electrodes for overall water splitting, J. Mater. Chem. A, 2018, 6, 14207–14214 RSC .
  125. Y. Y. Wu, G. D. Li, Y. P. Liu, L. Yang, X. R. Lian, T. Asefa and X. X. Zou, Overall Water Splitting Catalyzed Efficiently by an Ultrathin Nanosheet-Built, Hollow Ni3S2-Based Electrocatalyst, Adv. Funct. Mater., 2016, 26, 4839–4847 CrossRef CAS .
  126. X. Luo, Q. Zhou, S. Du, J. Li, J. Zhong, X. Deng and Y. Liu, Porous Co9S8/Nitrogen, Sulfur-Doped Carbon@Mo2C Dual Catalyst for Efficient Water Splitting, ACS Appl. Mater. Interfaces, 2018, 10, 22291–22302 CrossRef CAS .
  127. J. Y. Wang, T. Ouyang, Y. P. Deng, Y. S. Hong and Z. Q. Liu, Metallic Mo2C anchored pyrrolic-N induced N-CNTs/NiS2 for efficient overall water electrolysis, J. Power Sources, 2019, 420, 108–117 CrossRef CAS .
  128. J. J. Lv, J. Zhao, H. Fang, L. P. Jiang, L. L. Li, J. Ma and J. J. Zhu, Incorporating Nitrogen-Doped Graphene Quantum Dots and Ni3S2 Nanosheets: A Synergistic Electrocatalyst with Highly Enhanced Activity for Overall Water Splitting, Small, 2017, 13, 1700264 CrossRef .
  129. M. Guo, A. Qayum, S. Dong, X. Jiao, D. Chen and T. Wang, In situ conversion of metal (Ni, Co or Fe) foams into metal sulfide (Ni3S2, Co9S8 or FeS) foams with surface grown N-doped carbon nanotube arrays as efficient superaerophobic electrocatalysts for overall water splitting, J. Mater. Chem. A, 2020, 8, 9239–9247 RSC .
  130. H. Liu, C. Y. Xu, Y. Du, F. X. Ma, Y. Li, J. Yu and L. Zhen, Ultrathin Co9S8 nanosheets vertically aligned on N, S/rGO for low voltage electrolytic water in alkaline media, Sci. Rep., 2019, 9, 1951 CrossRef .
  131. Y. Tong, X. Yu and G. Shi, Cobalt disulfide/graphite foam composite films as self-standing electrocatalytic electrodes for overall water splitting, Phys. Chem. Chem. Phys., 2017, 19, 4821–4826 RSC .
  132. J. Hao, W. Yang, Z. Peng, C. Zhang, Z. Huang and W. Shi, A Nitrogen Doping Method for CoS2 Electrocatalysts with Enhanced Water Oxidation Performance, ACS Catal., 2017, 7, 4214–4220 CrossRef CAS .
  133. N. Yao, P. Li, Z. Zhou, R. Meng, G. Cheng and W. Luo, Nitrogen Engineering on 3D Dandelion-Flower-Like CoS2 for High-Performance Overall Water Splitting, Small, 2019, 15, 1901993 CrossRef .
  134. H. Liu, Q. He, H. Jiang, Y. Lin, Y. Zhang, M. Habib, S. Chen and L. Song, Electronic Structure Reconfiguration toward Pyrite NiS2 via Engineered Heteroatom Defect Boosting Overall Water Splitting, ACS Nano, 2017, 11, 11574–11583 CrossRef CAS .
  135. P. Chen, T. Zhou, M. Zhang, Y. Tong, C. Zhong, N. Zhang, L. Zhang, C. Wu and Y. Xie, 3D Nitrogen-Anion-Decorated Nickel Sulfides for Highly Efficient Overall Water Splitting, Adv. Mater., 2017, 29, 1701584 CrossRef .
  136. X. Y. Y. Lou and X. W. Lou, Mixed Metal Sulfides for Electrochemical Energy Storage and Conversion, Adv. Energy Mater., 2018, 8, 1701592 CrossRef .
  137. A. Sivanantham, P. Ganesan and S. Shanmugam, Hierarchical NiCo2S4 Nanowire Arrays Supported on Ni Foam: An Efficient and Durable Bifunctional Electrocatalyst for Oxygen and Hydrogen Evolution Reactions, Adv. Funct. Mater., 2016, 26, 4661–4672 CrossRef CAS .
  138. Y. Gong, H. Pan, Z. Xu, Z. Yang, Y. Lin and J. Wang, Crossed FeCo2S4 nanosheet arrays grown on 3D nickel foam as high-efficient electrocatalyst for overall water splitting, Int. J. Hydrogen Energy, 2018, 43, 17259–17264 CrossRef CAS .
  139. J. Yu, G. Cheng and W. Luo, Ternary nickel-iron sulfide microflowers as a robust electrocatalyst for bifunctional water splitting, J. Mater. Chem. A, 2017, 5, 15838–15844 RSC .
  140. D. Li, Z. Liu, J. Wang, B. Liu, Y. Qin, W. Yang and J. Liu, Hierarchical trimetallic sulfide FeCo2S4-NiCo2S4 nanosheet arrays supported on a Ti mesh: An efficient 3D bifunctional electrocatalyst for full water splitting, Electrochim. Acta, 2020, 340, 135957 CrossRef CAS .
  141. L. Wang, X. Duan, X. Liu, J. Gu, R. Si, Y. Qiu, Y. M. Qiu, D. Shi, F. Chen, X. Sun, J. Lin and J. Sun, Atomically Dispersed Mo Supported on Metallic Co9S8 Nanoflakes as an Advanced Noble-Metal-Free Bifunctional Water Splitting Catalyst Working in Universal pH Conditions, Adv. Energy Mater., 2020, 10, 1903137 CrossRef CAS .
  142. L. L. Feng, G. Yu, Y. Wu, G. D. Li, H. Li, Y. Sun, T. Asefa, W. Chen and X. Zou, High-index faceted Ni3S2 nanosheet arrays as highly active and ultrastable electrocatalysts for water splitting, J. Am. Chem. Soc., 2015, 137, 14023–14026 CrossRef CAS .
  143. Y. Li, Y. Wang, B. Pattengale, J. Yin, L. An, F. Cheng, Y. Li, J. Huang and P. Xi, High-index faceted CuFeS2 nanosheets with enhanced behavior for boosting hydrogen evolution reaction, Nanoscale, 2017, 9, 9230–9237 RSC .
  144. N. Zhang, J. Lei, J. Xie, H. Huang and Y. Yu, MoS2/Ni3S2 nanorod arrays well-aligned on Ni foam: a 3D hierarchical efficient bifunctional catalytic electrode for overall water splitting, RSC Adv., 2017, 7, 46286–46296 RSC .
  145. C. Guan, X. Liu, A. M. Elshahawy, H. Zhang, H. Wu, S. J. Pennycook and J. Wang, Metal-organic framework derived hollow CoS2 nanotube arrays: an efficient bifunctional electrocatalyst for overall water splitting, Nanoscale Horiz., 2017, 2, 342–348 RSC .
  146. C. Yan, J. Huang, C. Wu, Y. Li, Y. Tan, L. Zhang, Y. Sun, X. Huang and J. Xiong, In-situ formed NiS/Ni coupled interface for efficient oxygen evolution and hydrogen evolution, J. Mater. Sci. Technol., 2020, 42, 10–16 CrossRef .
  147. B. You, M. T. Tang, C. Tsai, F. A. Pedersen, X. Zheng and H. Li, Enhancing Electrocatalytic Water Splitting by Strain Engineering, Adv. Mater., 2019, 31, 1807001 CrossRef .
  148. X. Chen and G. Wang, Tuning the hydrogen evolution activity of MS2 (M = Mo or Nb) monolayers by strain engineering, Phys. Chem. Chem. Phys., 2016, 18, 9388–9395 RSC .
  149. Y. Shan and T. Li, Strain-regulated electronic structure for hydrogen evolution reaction in Fe3S4 monolayer, Phys. Lett. A, 2020, 384, 126368 CrossRef CAS .
  150. J. Zhang, W. Xiao, P. Xi, S. Xi, Y. Du, D. Gao and J. Ding, Activating and Optimizing Activity of CoS2 for Hydrogen Evolution Reaction through the Synergic Effect of N Dopants and S Vacancies, ACS Energy Lett., 2017, 2, 1022–1028 CrossRef CAS .
  151. J. Lin, P. Wang, H. Wang, C. Li, X. Si, J. Qi, J. Cao, Z. Zhong, W. Fei and J. Feng, Defect-Rich Heterogeneous MoS2/NiS2 Nanosheets Electrocatalysts for Efficient Overall Water Splitting, Adv. Sci., 2019, 6, 1900246 CrossRef .
  152. X. Zhu, J. Dai, L. Li, D. Zhao, Z. Wu, Z. Tang, L. J. Ma and S. Chen, Hierarchical carbon microflowers supported defect-rich Co3S4 nanoparticles: an efficient electrocatalyst for water splitting, Carbon, 2020, 160, 133–144 CrossRef CAS .
  153. Z. Cui, Y. Ge, H. Chu, R. Baines, P. Dong, J. Tang, Y. Yang, P. M. Ajayan, M. Ye and J. Shen, Controlled synthesis of Mo-doped Ni3S2 nanorods: an efficient and stable electro-catalyst for water splitting, J. Mater. Chem. A, 2017, 5, 1595–1602 RSC .
  154. G. Huang, Z. Xiao, R. Chen and S. Wang, Defect Engineering of Cobalt-Based Materials for Electrocatalytic Water Splitting, ACS Sustainable Chem. Eng., 2018, 6, 15954–15969 CrossRef CAS .
  155. S. Li, Y. Zhang and X. Niu, Defects and impurities induced structural and electronic changes in pyrite CoS2: first principles studies, Phys. Chem. Chem. Phys., 2018, 20, 11649–11655 RSC .
  156. W. Yang, J. Zeng, Y. Hua, C. Xu, S. S. Siwal and Q. Zhang, Defect engineering of cobalt microspheres by S doping and electrochemical oxidation as efficient bifunctional and durable electrocatalysts for water splitting at high current densities, J. Power Sources, 2019, 436, 226887 CrossRef CAS .
  157. X. Zhao, H. Liu, Y. Rao, X. Li, J. Wang, G. Xia and M. Wu, Carbon Dots Decorated Hierarchical NiCo2S4/Ni3S2 Composite for Efficient Water Splitting, ACS Sustainable Chem. Eng., 2019, 7, 2610–2618 CrossRef CAS .
  158. P. Hu, Z. Jia, H. Che, W. Zhou, N. Liu, F. Li and J. Wang, Engineering hybrid CoMoS4/Ni3S2 nanostructures as efficient bifunctional electrocatalyst for overall water splitting, J. Power Sources, 2019, 416, 95–103 CrossRef CAS .
  159. Y. Tan, M. Luo, P. Liu, C. Cheng, J. Han, K. Watanabe and M. Chen, Three-Dimensional Nanoporous Co9S4P4 Pentlandite as a Bifunctional Electrocatalyst for Overall Neutral Water Splitting, ACS Appl. Mater. Interfaces, 2019, 11, 3880–3888 CrossRef CAS .
  160. M. He, C. Feng, T. Liao, S. Hu, H. Wu and Z. Sun, Low-Cost Ni2P/Ni0.96S Heterostructured Bifunctional Electrocatalyst toward Highly Efficient Overall Urea-Water Electrolysis, ACS Appl. Mater. Interfaces, 2019, 12, 2225–2233 CrossRef .
  161. Y. Yang, K. Zhang, H. Lin, X. Li, H. C. Chan, L. Yang and Q. Gao, MoS2-Ni3S2 Heteronanorods as Efficient and Stable Bifunctional Electrocatalysts for Overall Water Splitting, ACS Catal., 2017, 7, 2357–2366 CrossRef CAS .
  162. F. Li, D. Zhang, R. C. Xu, W. F. Fu and X. J. Lv, Superhydrophilic Heteroporous MoS2/Ni3S2 for Highly Efficient Electrocatalytic Overall Water Splitting, ACS Appl. Energy Mater., 2018, 1, 3929–3936 CrossRef CAS .
  163. Q. Qin, L. Chen, T. Wei and X. Liu, MoS2/NiS Yolk-Shell Microsphere-Based Electrodes for Overall Water Splitting and Asymmetric Supercapacitor, Small, 2019, 15, 1803639 CrossRef .
  164. D. Zhang, L. Jiang, Y. Liu, L. Qiu, J. Zhang and D. Yuan, Ni3S2-MoSx nanorods grown on Ni foam as high-efficient electrocatalysts for overall water splitting, Int. J. Hydrogen Energy, 2019, 44, 17900–17908 CrossRef CAS .
  165. N. Huang, S. Yan, M. Zhang, Y. Ding, L. Yang, P. Sun and X. Sun, A MoS2-Co9S8-NC heterostructure as an efficient bifunctional electrocatalyst towards hydrogen and oxygen evolution reaction, Electrochim. Acta, 2019, 327, 134942 CrossRef CAS .
  166. S. Shit, S. Chhetri, S. Bolar, N. C. Murmu, W. Jang, H. Koo and T. Kuila, Hierarchical Cobalt Sulfide/Molybdenum Sulfide Heterostructure as Bifunctional Electrocatalyst towards Overall Water Splitting, ChemElectroChem, 2019, 6, 430–438 CrossRef CAS .
  167. M. S. Islam, M. Kim, X. Jin, S. M. Oh, N. S. Lee, H. Kim and S. J. Hwang, Bifunctional 2D Superlattice Electrocatalysts of Layered Double Hydroxide-Transition Metal Dichalcogenide Active for Overall Water Splitting, ACS Energy Lett., 2018, 3, 952–960 CrossRef CAS .
  168. Y. Liu, S. Jiang, S. Li, L. Zhou, Z. Li, J. Li and M. Shao, Interface engineering of (Ni, Fe)S2@MoS2 heterostructures for synergetic electrochemical water splitting, Appl. Catal. B Environ., 2019, 247, 107–114 CrossRef CAS .
  169. P. Kuang, M. He, H. Zou, J. Yu and K. Fan, 0D/3D MoS2-NiS2/N-doped graphene foam composite for efficient overall water splitting, Appl. Catal. B Environ., 2019, 254, 15–25 CrossRef CAS .
  170. T. Yoon and K. S. Kim, One-Step Synthesis of CoS-Doped β-Co(OH)2@Amorphous MoS2+x Hybrid Catalyst Grown on Nickel Foam for High-Performance Electrochemical Overall Water Splitting, Adv. Funct. Mater., 2016, 26, 7386–7393 CrossRef CAS .
  171. Y. Zhu, L. Song, N. Song, M. Li, C. Wang and X. Lu, Bifunctional and Efficient CoS2-C@MoS2 Core-Shell Nanofiber Electrocatalyst for Water Splitting, ACS Sustainable Chem. Eng., 2019, 7, 2899–2905 CrossRef CAS .
  172. V. Ganesan and J. Kim, Multi-shelled CoS2-MoS2 hollow spheres as efficient bifunctional electrocatalysts for overall water splitting, Int. J. Hydrogen Energy, 2020, 45, 13290–13299 CrossRef CAS .
  173. Y. Guo, J. Tang, Z. Wang, Y. M. Kang, Y. Bando and Y. Yamauchi, Elaborately assembled core-shell structured metal sulfides as a bifunctional catalyst for highly efficient electrochemical overall water splitting, Nano Energy, 2018, 47, 494–502 CrossRef CAS .
  174. A. Muthurasu, V. Maruthapandian and H. Y. Kim, Metal-organic framework derived Co3O4/MoS2 heterostructure for efficient bifunctional electrocatalysts for oxygen evolution reaction and hydrogen evolution reaction, Appl. Catal. B Environ., 2019, 248, 202–210 CrossRef CAS .
  175. Y. Wang, T. Williams, T. Gengenbach, B. Kong, D. Zhao, H. Wang and C. Selomulya, Unique hybrid Ni2P/MoO2@MoS2 nanomaterials as bifunctional non-noble-metal electro-catalysts for water splitting, Nanoscale, 2017, 9, 17349–17356 RSC .
  176. R. Prasannachandran, T. V. Vineesh, M. B. Lithin, R. Nandakishore and M. M. Shaijumon, Phosphorene-quantum-dot-interspersed few-layered MoS2 hybrids as efficient bifunctional electrocatalysts for hydrogen and oxygen evolution, Chem. Commun., 2020, 56, 8623–8626 RSC .
  177. Y. Xu, X. Chai, T. Ren, H. Yu, S. Yin, Z. Wang, X. Li, L. Wang and H. Wang, Synergism of Interface and Electronic Effects: Bifunctional N-Doped Ni3S2/N-Doped MoS2 Hetero-Nanowires for Efficient Electrocatalytic Overall Water Splitting, Chem.–Eur J., 2019, 25, 16074–16080 CrossRef CAS .
  178. Z. Zhai, C. Li, L. Zhang, H. C. Wu, L. Zhang, N. Tang, W. Wang and J. Gong, Dimensional construction and morphological tuning of heterogeneous MoS2/NiS electrocatalysts for efficient overall water splitting, J. Mater. Chem. A, 2018, 6, 9833–9838 RSC .
  179. A. Muthurasu, G. P. Ojha, M. Lee and H. Y. Kim, Zeolitic imidazolate framework derived Co3S4 hybridized MoS2-Ni3S2 heterointerface for electrochemical overall water splitting reactions, Electrochim. Acta, 2020, 334, 135537 CrossRef CAS .
  180. C. Qin, A. Fan, X. Zhang, S. Wang, X. Yuan and X. Dai, Interface engineering: few-layer MoS2 coupled to a NiCo-sulfide nanosheet heterostructure as a bifunctional electrocatalyst for overall water splitting, J. Mater. Chem. A, 2019, 7, 27594–27602 RSC .
  181. Y. Wu, F. Li, W. Chen, Q. Xiang, Y. Ma, H. Zhu, P. Tao, C. Song, W. Shang, T. Deng and J. Wu, Coupling Interface Constructions of MoS2/Fe5Ni4S8 Heterostructures for Efficient Electrochemical Water Splitting, Adv. Mater., 2018, 30, 1803151 CrossRef .
  182. H. Li, S. Chen, X. Jia, B. Xu, H. Lin, H. Yang, L. Song and X. Wang, Amorphous nickel-cobalt complexes hybridized with 1T-phase molybdenum disulfide via hydrazine-induced phase transformation for water splitting, Nat. Commun., 2017, 8, 15377 CrossRef CAS .
  183. T. Yang, L. Yin, M. He, W. Wei, G. Cao, X. Ding, Y. Wang, Z. Zhao, T. Yu, H. Zhao and D. Zhang, Yolk-shell hierarchical catalyst with tremella-like molybdenum sulfide on transition metal (Co, Ni and Fe) sulfide for electrochemical water splitting, Chem. Commun., 2019, 55, 14343–14346 RSC .
  184. J. Y. Xue, F. L. Li, Z. Y. Zhao, C. Li, C. Y. Ni, H. W. Gu, P. Braunstein, X. Q. Huang and J. P. Lang, A hierarchically-assembled Fe-MoS2/Ni3S2/nickel foam electrocatalyst for efficient water splitting, Dalton Trans., 2019, 48, 12186–12192 RSC .
  185. X. Zou, Y. Wu, Y. Liu, D. Liu, W. Li, L. Gu, H. Liu, P. Wang, L. Sun and Y. Zhang, In Situ Generation of Bifunctional, Efficient Fe-Based Catalysts from Mackinawite Iron Sulfide for Water Splitting, Chem, 2018, 4, 1139–1152 CAS .
  186. S. Shit, S. Bolar, N. C. Murmu and T. Kuila, Binder-Free Growth of Nickel-Doped Iron Sulfide on Nickel Foam via Electrochemical Deposition for Electrocatalytic Water Splitting, ACS Sustainable Chem. Eng., 2019, 7, 18015–18026 CrossRef CAS .
  187. L. Gao, C. Guo, X. Liu, X. Ma, M. Zhao, X. Kuang, H. Yang, X. Zhu, X. Sun and Q. Wei, Co-Doped FeS2 with a porous structure for efficient electrocatalytic overall water splitting, New J. Chem., 2020, 44, 1711–1718 RSC .
  188. S. Wang, P. Ning, S. Huang, W. Wang, S. Fei, Q. He, J. Zai, Y. Jiang, Z. Hu, X. Qian and Z. Chen, Multi-functional NiS2/FeS2/N-doped carbon nanorods derived from metal-organic frameworks with fast reaction kinetics for high performance overall water splitting and lithium-ion batteries, J. Power Sources, 2019, 436, 226857 CrossRef CAS .
  189. R. Zhang, Z. Zhu, J. Lin, K. Zhang, N. Li and C. Zhao, Hydrolysis assisted in situ growth of 3D hierarchical FeS/NiS/nickel foam electrode for overall water splitting, Electrochim. Acta, 2020, 332, 135534 CrossRef CAS .
  190. N. Yao, T. Tan, F. Yang, G. Cheng and W. Luo, Well-aligned metal-organic framework array-derived CoS2 nanosheets toward robust electrochemical water splitting, Mater. Chem. Front., 2018, 2, 1732–1738 RSC .
  191. R. Souleymen, Z. T. Wang, C. Qiao, M. Naveed and C. B. Cao, Microwave-assisted synthesis of graphene-like cobalt sulfide freestanding sheets as an efficient bifunctional electrocatalyst for overall water splitting, J. Mater. Chem. A, 2018, 6, 7592–7607 RSC .
  192. Y. Li, X. Fu, W. Zhu, J. Gong, J. Sun, D. Zhang and J. Wang, Self-ZIF template-directed synthesis of a CoS nanoflake array as a Janus electrocatalyst for overall water splitting, Inorg. Chem. Front., 2019, 6, 2090–2095 RSC .
  193. Y. Yang, M. Yuan, H. Li, G. Sun and S. Ma, Controllable synthesis of ultrathin Co9S8 nanosheets as a highly efficient electrocatalyst for overall water splitting, Electrochim. Acta, 2018, 281, 198–207 CrossRef CAS .
  194. M. Zhu, Z. Zhang, H. Zhang, H. Zhang, X. Zhang, L. Zhang and S. Wang, Hydrophilic cobalt sulfide nanosheets as a bifunctional catalyst for oxygen and hydrogen evolution in electrolysis of alkaline aqueous solution, J. Colloid Interface Sci., 2018, 509, 522–528 CrossRef CAS .
  195. X. Du, G. Ma and X. Zhang, Mo-doped Co9S8 nanorod array as a high performance electrochemical water splitting catalyst in alkaline solution, Int. J. Hydrogen Energy, 2019, 44, 27765–27771 CrossRef CAS .
  196. X. Du, H. Su and X. Zhang, Metal-Organic Framework-Derived Cu-Doped Co9S8 Nanorod Array with Less Low-Valence Co Sites as Highly Efficient Bifunctional Electrodes for Overall Water Splitting, ACS Sustainable Chem. Eng., 2019, 7, 16917–16926 CrossRef CAS .
  197. X. Du, H. Su and X. Zhang, Cr doped-Co9S8 nanoarrays as high-efficiency electrocatalysts for water splitting, J. Alloys Compd., 2020, 824, 153965 CrossRef CAS .
  198. X. Du, H. Su and X. Zhang, Metal-organic framework-derived M (M = Fe, Ni, Zn and Mo) doped Co9S8 nanoarrays as efficient electrocatalyst for water splitting: The combination of theoretical calculation and experiment, J. Catal., 2020, 383, 103–116 CrossRef CAS .
  199. S. Tang, X. Wang, Y. Zhang, M. Courte, H. J. Fan and D. Fichou, Combining Co3S4 and Ni:Co3S4 nanowires as efficient catalysts for overall water splitting: an experimental and theoretical study, Nanoscale, 2019, 11, 2202–2210 RSC .
  200. L. Lei, D. Huang, C. Zhang, R. Deng, S. Chen and Z. Li, F dopants triggered active sites in bifunctional cobalt sulfide@nickel foam toward electrocatalytic overall water splitting in neutral and alkaline media: Experiments and theoretical calculations, J. Catal., 2020, 385, 129–139 CrossRef CAS .
  201. Y. Li, Z. Mao, Q. Wang, D. Li, R. Wang, B. He, Y. Gong and H. Wang, Hollow nanosheet array of phosphorus-anion-decorated cobalt disulfide as an efficient electrocatalyst for overall water splitting, Chem. Eng. J., 2020, 390, 124556 CrossRef CAS .
  202. Z. Dai, H. Geng, J. Wang, Y. Luo, B. Li, Y. Zong, J. Yang, Y. Guo, Y. Zheng, X. Wang and Q. Yan, Hexagonal-Phase Cobalt Monophosphosulfide for Highly Efficient Overall Water Splitting, ACS Nano, 2017, 11, 11031–11040 CrossRef CAS .
  203. S. Deng, Y. Zhong, Y. Zeng, Y. Wang, X. Wang, X. Lu, X. Xia and J. Tu, Hollow TiO2@Co9S8 Core-Branch Arrays as Bifunctional Electrocatalysts for Efficient Oxygen/Hdrogen Production, Adv. Sci., 2018, 5, 1700772 CrossRef .
  204. C. Ray, S. C. Lee, K. V. Sankar, B. Jin, J. Lee, J. H. Park and S. C. Jun, Amorphous Phosphorus-Incorporated Cobalt Molybdenum Sulfide on Carbon Cloth: An Efficient and Stable Electrocatalyst for Enhanced Overall Water Splitting over Entire pH Values, ACS Appl. Mater. Interfaces, 2017, 9, 37739–37749 CrossRef CAS .
  205. Y. Chen, S. Xu, S. Zhu, R. J. Jacob, G. Pastel, Y. Wang, Y. Li, J. Dai, F. Chen, H. Xie, B. Liu, Y. Yao, L. G. S. Riba, M. R. Zachariah, T. Li and L. Hu, Millisecond synthesis of CoS nanoparticles for highly efficient overall water splitting, Nano Res., 2019, 12, 2259–2267 CrossRef CAS .
  206. W. Zhang, X. Ma, C. Zhong, T. Ma, Y. Deng, W. Hu and X. Han, Pyrite-Type CoS2 Nanoparticles Supported on Nitrogen-Doped Graphene for Enhanced Water Splitting, Front. Chem., 2018, 6, 569 CrossRef CAS .
  207. S. Huang, Y. Meng, S. He, A. Goswami, Q. Wu, J. Li, S. Tong, T. Asefa and M. Wu, N-, O-, and S-Tridoped Carbon-Encapsulated Co9S8 Nanomaterials: Efficient Bifunctional Electrocatalysts for Overall Water Splitting, Adv. Funct. Mater., 2017, 27, 1606585 CrossRef .
  208. J. Y. Wang, T. Ouyang, N. Li, T. Y. Ma and Z. Q. Liu, 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, 1130–1140 CrossRef CAS .
  209. J. Li, W. Xu, J. Luo, D. Zhou, D. Zhang, L. Wei, P. Xu and D. Yuan, Synthesis of 3D Hexagram-Like Cobalt-Manganese Sulfides Nanosheets Grown on Nickel Foam: A Bifunctional Electrocatalyst for Overall Water Splitting, Nanomicro Lett., 2018, 10, 6 CrossRef .
  210. R. Huang, W. Chen, Y. Zhang, Z. Huang, H. Dai, Y. Zhou, Y. Wu and X. Lv, Well-designed cobalt-nickel sulfide microspheres with unique peapod-like structure for overall water splitting, J. Colloid Interface Sci., 2019, 556, 401–410 CrossRef CAS .
  211. Z. Liang, Z. Yang, J. Dang, J. Qi, H. Yuan, J. Gao, W. Zhang, H. Zheng and R. Cao, Hollow Bimetallic Zinc Cobalt Phosphosulfides for Efficient Overall Water Splitting, Chemistry, 2019, 25, 621–626 CAS .
  212. S. Czioska, J. Wang, X. Teng and Z. Chen, Hierarchically Structured CuCo2S4 Nanowire Arrays as Efficient Bifunctional Electrocatalyst for Overall Water Splitting, ACS Sustainable Chem. Eng., 2018, 6, 11877–11883 CrossRef CAS .
  213. C. Zequine, S. Bhoyate, F. Wang, X. Li, K. Siam, P. K. Kahol and R. K. Gupta, Effect of solvent for tailoring the nanomorphology of multinary CuCo2S4 for overall water splitting and energy storage, J. Alloys Compd., 2019, 784, 1–7 CrossRef CAS .
  214. X. Du, J. Fu and X. Zhang, Controlled Synthesis of CuCo2S4@Ni(OH)2 Hybrid Nanorod Arrays for Water Splitting at an Ultralow Cell Voltage of 1.47 V, Chem. Asian J., 2019, 14, 3386–3396 CrossRef CAS .
  215. J. Hu, Y. Ou, Y. Li, D. Gao, Y. Zhang and P. Xiao, FeCo2S4 Nanosheet Arrays Supported on Ni Foam: An Efficient and Durable Bifunctional Electrocatalyst for Overall Water-Splitting, ACS Sustainable Chem. Eng., 2018, 6, 11724–11733 CrossRef CAS .
  216. L. Hui, Y. Xue, D. Jia, Z. Zuo, Y. Li, H. Liu, Y. Zhao and Y. Li, Controlled Synthesis of a Three-Segment Heterostructure for High-Performance Overall Water Splitting, ACS Appl. Mater. Interfaces, 2018, 10, 1771–1780 CrossRef CAS .
  217. W. Zhu, M. Ren, N. Hu, W. Zhang, Z. Luo, R. Wang, J. Wang, L. Huang, Y. Suo and J. Wang, Traditional NiCo2S4 Phase with Porous Nanosheets Array Topology on Carbon Cloth: A Flexible, Versatile and Fabulous Electrocatalyst for Overall Water and Urea Electrolysis, ACS Sustainable Chem. Eng., 2018, 6, 5011–5020 CrossRef CAS .
  218. J. Yu, C. Lv, L. Zhao, L. Zhang, Z. Wang and Q. Liu, Reverse Microemulsion-Assisted Synthesis of NiCo2S4 Nanoflakes Supported on Nickel Foam for Electrochemical Overall Water Splitting, Adv. Mater. Interfaces, 2018, 5, 701396 Search PubMed .
  219. Y. Gong, Y. Lin, Z. Yang, J. Wang, H. Pan, Z. Xu and Y. Liu, Crossed NiCo2S4 Nanowires Supported on Nickel Foam as a Bifunctional Catalyst for Efficient Overall Water Splitting, ChemistrySelect, 2019, 4, 1180–1187 CrossRef CAS .
  220. Y. Tang, H. Yang, J. Sun, M. Xia, W. Guo, L. Yu, J. Yan, J. Zheng, L. Chang and F. Gao, Phase-pure pentlandite Ni4.3Co4.7S8 binary sulfide as an efficient bifunctional electrocatalyst for oxygen evolution and hydrogen evolution, Nanoscale, 2018, 10, 10459–10466 RSC .
  221. Q. Zhang, C. Ye, X. L. Li, Y. H. Deng, B. X. Tao, W. Xiao, L. J. Li, N. B. Li and H. Q. Luo, Self-Interconnected Porous Networks of NiCo Disulfide as Efficient Bifunctional Electrocatalysts for Overall Water Splitting, ACS Appl. Mater. Interfaces, 2018, 10, 27723–27733 CrossRef CAS .
  222. Z. Ma, H. Meng, M. Wang, B. Tang, J. Li and X. Wang, Porous Ni-Mo-S Nanowire Network Film Electrode as a High-Efficiency Bifunctional Electrocatalyst for Overall Water Splitting, ChemElectroChem, 2018, 5, 335–342 CrossRef CAS .
  223. K. L. Yan, J. F. Qin, Z. Z. Liu, B. Dong, J. Q. Chi, W. K. Gao, J. H. Lin, Y. M. Chai and C. G. Liu, 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 CAS .
  224. H. Fan, J. Huang, G. Chen, W. Chen, R. Zhang, S. Chu, X. Wang, C. Li and K. K. Ostrikov, Hollow Ni-V-Mo Chalcogenide Nanopetals as Bifunctional Electrocatalyst for Overall Water Splitting, ACS Sustainable Chem. Eng., 2019, 7, 1622–1632 CrossRef CAS .
  225. Y. Ning, D. Ma, Y. Shen, F. Wang and X. Zhang, Constructing hierarchical mushroom-like bifunctional NiCo/NiCo2S4@NiCo/Ni foam electrocatalysts for efficient overall water splitting in alkaline media, Electrochim. Acta, 2018, 265, 19–31 CrossRef CAS .
  226. Y. Liu, Q. Li, R. Si, G. D. Li, W. Li, D. P. Liu, D. Wang, L. Sun, Y. Zhang and X. Zou, Coupling Sub-Nanometric Copper Clusters with Quasi-Amorphous Cobalt Sulfide Yields Efficient and Robust Electrocatalysts for Water Splitting Reaction, Adv. Mater., 2017, 29, 1606200 CrossRef .
  227. G. B. Darband, M. Aliofkhazraei, S. Hyun, A. S. Rouhaghdam and S. Shanmugam, Electrodeposition of Ni-Co-Fe mixed sulfide ultrathin nanosheets on Ni nanocones: a low-cost, durable and high performance catalyst for electrochemical water splitting, Nanoscale, 2019, 11, 16621–16634 RSC .
  228. W. Chen, Y. Zhang, G. Chen, R. Huang, Y. Wu, Y. Zhou, Y. Hu and K. K. Ostrikov, Hierarchical porous bimetal-sulfide bi-functional nanocatalysts for hydrogen production by overall water electrolysis, J. Colloid Interface Sci., 2020, 560, 426–435 CrossRef CAS .
  229. C. Liu, D. Jia, Q. Hao, X. Zheng, Y. Li, C. Tang, H. Liu, J. Zhang and X. Zheng, P-Doped Iron-Nickel Sulfide Nanosheet Arrays for Highly Efficient Overall Water Splitting, ACS Appl. Mater. Interfaces, 2019, 11, 27667–27676 CrossRef CAS .
  230. Q. Che, Q. Li, Y. Tan, X. Chen, X. Xu and Y. Chen, One-step controllable synthesis of amorphous (Ni-Fe)S/NiFe(OH) hollow microtube/sphere films as superior bifunctional electrocatalysts for quasi-industrial water splitting at large-current-density, Appl. Catal. B Environ., 2019, 246, 337–348 CrossRef CAS .
  231. F. Wu, X. Guo, G. Hao, Y. Hu and W. Jiang, Self-supported hollow Co(OH)2/NiCo sulfide hybrid nanotube arrays as efficient electrocatalysts for overall water splitting, J. Solid State Electrochem., 2019, 23, 2627–2637 CrossRef CAS .
  232. Y. Hou, M. Qiu, G. Nam, M. G. Kim, T. Zhang, K. Liu, X. Zhuang, J. Cho, C. Yuan and X. Feng, Integrated Hierarchical Cobalt Sulfide/Nickel Selenide Hybrid Nanosheets as an Efficient Three-dimensional Electrode for Electrochemical and Photoelectrochemical Water Splitting, Nano Lett., 2017, 17, 4202–4209 CrossRef CAS .
  233. T. Xiong, G. Li, D. J. Young, Z. Tan, X. H. Yin, Y. Mi and F. Hu, In-situ surface-derivation of Ni-Mo bimetal sulfides nanosheets on Co3O4 nanoarrays as an advanced overall water splitting electrocatalyst in alkaline solution, J. Alloys Compd., 2019, 791, 328–335 CrossRef CAS .
  234. F. Li, R. Xu, Y. Li, F. Liang, D. Zhang, W. F. Fu and X. J. Lv, N-doped carbon coated NiCo2S4 hollow nanotube as bifunctional electrocatalyst for overall water splitting, Carbon, 2019, 145, 521–528 CrossRef CAS .
  235. R. Miao, J. He, S. Sahoo, Z. Luo, W. Zhong, S. Y. Chen, C. Guild, T. Jafari, B. Dutta, S. A. Cetegen, M. Wang, S. P. Alpay and S. L. Suib, Reduced Graphene Oxide Supported Nickel-Manganese-Cobalt Spinel Ternary Oxide Nanocomposites and Their Chemically Converted Sulfide Nanocomposites as Efficient Electrocatalysts for Alkaline Water Splitting, ACS Catal., 2017, 7, 819–832 CrossRef CAS .
  236. B. Wu, H. Qian, Z. Nie, Z. Luo, Z. Wu, P. Liu, H. He, J. Wu, S. Chen and F. Zhang, Ni3S2 nanorods growing directly on Ni foam for all-solid-state asymmetric supercapacitor and efficient overall water splitting, J. Energy Chem., 2020, 46, 178–186 CrossRef .
  237. T. Zhu, L. Zhu, J. Wang and G. W. Ho, In situ chemical etching of tunable 3D Ni3S2 superstructures for bifunctional electrocatalysts for overall water splitting, J. Mater. Chem. A, 2016, 4, 13916–13922 RSC .
  238. G. Ren, Q. Hao, J. Mao, L. Liang, H. Liu, C. Liu and J. Zhang, Ultrafast fabrication of nickel sulfide film on Ni foam for efficient overall water splitting, Nanoscale, 2018, 10, 17347–17353 RSC .
  239. H. Ren, Z. H. Huang, Z. Yang, S. Tang, F. Kang and R. Lv, Facile synthesis of free-standing nickel chalcogenide electrodes for overall water splitting, J. Energy Chem., 2017, 26, 1217–1222 CrossRef .
  240. Y. Guo, D. Guo, F. Ye, K. Wang and Z. Shi, Synthesis of lawn-like NiS2 nanowires on carbon fiber paper as bifunctional electrode for water splitting, Int. J. Hydrogen Energy, 2017, 42, 17038–17048 CrossRef CAS .
  241. A. P. Tiwari, Y. Yoon, T. G. Novak, K. S. An and S. Jeon, Continuous Network of Phase-Tuned Nickel Sulfide Nanostructures for Electrocatalytic Water Splitting, ACS Appl. Nano Mater., 2019, 2, 5061–5070 CrossRef CAS .
  242. J. Ding, S. Ji, H. Wang, H. Gai, F. Liu, V. Linkov and R. Wang, Mesoporous nickel-sulfide/nickel/N-doped carbon as HER and OER bifunctional electrocatalyst for water electrolysis, Int. J. Hydrogen Energy, 2019, 44, 2832–2840 CrossRef CAS .
  243. L. Zeng, K. Sun, Y. Chen, Z. Liu, Y. Chen, Y. Pan, R. Zhao, Y. Liu and C. Liu, Neutral-pH overall water splitting catalyzed efficiently by a hollow and porous structured ternary nickel sulfoselenide electrocatalyst, J. Mater. Chem. A, 2019, 7, 16793–16802 RSC .
  244. J. Jian, L. Yuan, H. Qi, X. Sun, L. Zhang, H. Li, H. Yuan and S. Feng, Sn-Ni3S2 Ultrathin Nanosheets as Efficient Bifunctional Water-Splitting Catalysts with a Large Current Density and Low Overpotential, ACS Appl. Mater. Interfaces, 2018, 10, 40568–40576 CrossRef CAS .
  245. C. Wu, B. Liu, J. Wang, Y. Su, H. Yan, C. Ng, C. Li and J. Wei, 3D structured Mo-doped Ni3S2 nanosheets as efficient dual-electrocatalyst for overall water splitting, Appl. Surf. Sci., 2018, 441, 1024–1033 CrossRef CAS .
  246. Q. Jia, X. Wang, S. Wei, C. Zhou, J. Wang and J. Liu, Porous flower-like Mo-doped NiS heterostructure as highly efficient and robust electrocatalyst for overall water splitting, Appl. Surf. Sci., 2019, 484, 1052–1060 CrossRef CAS .
  247. Y. Gong, Z. Yang, Y. Zhi, Y. Lin, T. Zhou, J. Li, F. Jiao and W. Wang, Controlled synthesis of bifunctional particle-like Mo/Mn-NixSy/NF electrocatalyst for highly efficient overall water splitting, Dalton Trans., 2019, 48, 6718–6729 RSC .
  248. X. Wang, W. Zhang, J. Zhang and Z. Wu, Fe-Doped Ni3S2 Nanowires with Surface-Restricted Oxidation Toward High-Current-Density Overall Water Splitting, ChemElectroChem, 2019, 6, 4550–4559 CrossRef CAS .
  249. W. Zhu, Z. Yue, W. Zhang, N. Hu, Z. Luo, M. Ren, Z. Xu, Z. Wei, Y. Suo and J. Wang, Wet-chemistry topotactic synthesis of bimetallic iron-nickel sulfide nanoarrays: an advanced and versatile catalyst for energy efficient overall water and urea electrolysis, J. Mater. Chem. A, 2018, 6, 4346–4353 RSC .
  250. S. Song, Y. Wang, W. Li, P. Tian, S. Zhou, H. Gao, X. Tian and J. Zang, Co-doped Ni3S2 hierarchical nanoarrays derived from zeolitic imidazolate frameworks as bifunctional electrocatalysts for highly enhanced overall-water-splitting activity, J. Alloys Compd., 2020, 827, 154299 CrossRef CAS .
  251. J. Guo, K. Zhang, Y. Sun, Q. Liu, L. Tang and X. Zhang, Efficient bifunctional vanadium-doped Ni3S2 nanorod array for overall water splitting, Inorg. Chem. Front., 2019, 6, 443–450 RSC .
  252. Q. Liu, J. Huang, L. Cao, K. Kajiyoshi, K. Li, Y. Feng, C. Fu, L. Kou and L. Feng, V-Doping Triggered Formation and Structural Evolution of Dendritic Ni3S2@NiO Core-Shell Nanoarrays for Accelerating Alkaline Water Splitting, ACS Sustainable Chem. Eng., 2020, 8, 6222–6233 CrossRef CAS .
  253. S. Deng, K. Zhang, D. Xie, Y. Zhang, Y. Zhang, Y. Wang, J. Wu, X. Wang, H. J. Fan, X. Xia and J. Tu, High-Index-Faceted Ni3S2 Branch Arrays as Bifunctional Electrocatalysts for Efficient Water Splitting, Nanomicro Lett., 2019, 11, 12 CAS .
  254. Q. Lv, L. Yang, W. Wang, S. Lu, T. Wang, L. Cao and B. Dong, One-step construction of core/shell nanoarrays with a holey shell and exposed interfaces for overall water splitting, J. Mater. Chem. A, 2019, 7, 1196–1205 RSC .
  255. Y. Zhao, B. Jin, A. Vasileff, Y. Jiao and S. Z. Qiao, Interfacial nickel nitride/sulfide as a bifunctional electrode for highly efficient overall water/seawater electrolysis, J. Mater. Chem. A, 2019, 7, 8117–8121 RSC .
  256. Y. Zhang, J. Fu, H. Zhao, R. Jiang, F. Tian and R. Zhang, Tremella-like Ni3S2/MnS with ultrathin nanosheets and abundant oxygen vacancies directly used for high speed overall water splitting, Appl. Catal. B Environ., 2019, 257, 117899 CrossRef CAS .
  257. J. Zhu, M. Sun, S. Liu, X. Liu, K. Hu and L. Wang, Study of active sites on Se-MnS/NiS heterojunctions as highly efficient bifunctional electrocatalysts for overall water splitting, J. Mater. Chem. A, 2019, 7, 26975–26983 RSC .
  258. S. Qu, J. Huang, J. Yu, G. Chen, W. Hu, M. Yin, R. Zhang, S. Chu and C. Li, Ni3S2 Nanosheet Flowers Decorated with CdS Quantum Dots as a Highly Active Electrocatalysis Electrode for Synergistic Water Splitting, ACS Appl. Mater. Interfaces, 2017, 9, 29660–29668 CrossRef CAS .
  259. Y. Wu, Y. Liu, G. D. Li, X. Zou, X. Lian, D. Wang, L. Sun, T. Asefa and X. Zou, Efficient electrocatalysis of overall water splitting by ultrasmall NixCo3−xS4 coupled Ni3S2 nanosheet arrays, Nano Energy, 2017, 35, 161–170 CrossRef CAS .
  260. L. An, J. Feng, Y. Zhang, R. Wang, H. Liu, G. C. Wang, F. Cheng and P. Xi, Epitaxial Heterogeneous Interfaces on N-NiMoO4/NiS2 Nanowires/Nanosheets to Boost Hydrogen and Oxygen Production for Overall Water Splitting, Adv. Funct. Mater., 2019, 29, 1805298 CrossRef .
  261. M. Zheng, J. Du, B. Hou and C. L. Xu, Few-Layered Mo(1-x)WxS2 Hollow Nanospheres on Ni3S2 Nanorod Heterostructure as Robust Electrocatalysts for Overall Water Splitting, ACS Appl. Mater. Interfaces, 2017, 9, 26066–26076 CrossRef CAS .
  262. Z. Ma, Q. Zhao, J. Li, B. Tang, Z. Zhang and X. Wang, 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 .
  263. F. Du, L. Shi, Y. Zhang, T. Li, J. Wang, G. Wen, A. Alsaedi, T. Hayat, Y. Zhou and Z. Zou, Foam-like Co9S8/Ni3S2 heterostructure nanowire arrays for efficient bifunctional overall water-splitting, Appl. Catal. B Environ., 2019, 253, 246–252 CrossRef CAS .
  264. X. Du, C. Huang and X. Zhang, Surface modification of a Co9S8 nanorods with Ni(OH)2 on nickel foam for high water splitting performance, Int. J. Hydrogen Energy, 2019, 44, 19953–19966 CrossRef CAS .
  265. S. Shit, S. Chhetri, W. Jang, N. C. Murmu, H. Koo, P. Samanta and T. Kuila, 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, 27712–27722 CrossRef CAS .
  266. S. Q. Qua, W. Chen, J. S. Yu, G. L. Chen, R. Zhang, S. J. Chu, J. Huang, X. Q. Wang, C. R. Li and K. Ostrikovd, Cross-linked trimetallic nanopetals for electrocatalytic water splitting, J. Power Sources, 2018, 390, 224–230 CrossRef .
  267. K. S. Bhat and H. S. Nagaraja, In Situ Synthesis of Copper Sulfide-Nickel Sulfide Arrays on Three-Dimensional Nickel Foam for Overall Water Splitting, ChemistrySelect, 2020, 5, 2455–2464 CrossRef CAS .
  268. D. Yang, L. Cao, L. Feng, J. Huang, K. Kajiyoshi, Y. Feng, Q. Liu, W. Li, L. Feng and G. Hai, Formation of hierarchical Ni3S2 nanohorn arrays driven by in situ generation of VS4 nanocrystals for boosting alkaline water splitting, Appl. Catal. B Environ., 2019, 257, 117911 CrossRef CAS .
  269. Y. Q. Gong, Y. Lin, Z. Yang, F. X. Jiao, J. H. Li and W. F. Wang, High-performance bifunctional flower-like Mn-doped Cu7.2S4@NiS2@NiS/NF catalyst for overall water splitting, Appl. Surf. Sci., 2019, 476, 840–849 CrossRef CAS .
  270. X. Luo, P. Ji, P. Wang, R. Cheng, D. Chen, C. Lin, J. Zhang, J. He, Z. Shi, N. Li, S. Xiao and S. Mu, Interface Engineering of Hierarchical Branched Mo-Doped Ni3S2/NixPy Hollow Heterostructure Nanorods for Efficient Overall Water Splitting, Adv. Energy Mater., 2020, 10, 1903891 CrossRef CAS .
  271. L. Zeng, K. Sun, X. Wang, Y. Liu, Y. Pan, Z. Liu, D. Cao, Y. Song, S. Liu and C. Liu, Three-dimensional-networked Ni2P/Ni3S2 heteronanoflake arrays for highly enhanced electrochemical overall-water-splitting activity, Nano Energy, 2018, 51, 26–36 CrossRef CAS .
  272. X. Xiao, D. K. Huang, Y. Q. Fu, M. Wen, X. X. Jiang, X. W. Lv, M. Li, L. Gao, S. S. Liu, M. K. Wang, C. Zhao and Y. Shen, Engineering NiS/Ni2P Heterostructures for Efficient Electrocatalytic Water Splitting, ACS Appl. Mater. Interfaces, 2018, 10, 4689–4696 CrossRef CAS .
  273. X. W. Zhong, J. Tang, J. W. Wang, M. M. Shao, J. W. Chai, S. P. Wang, M. Yang, Y. Yang, N. Wang, S. J. Wang, B. M. Xu and H. Pan, 3D heterostructured pure and N-Doped Ni3S2/VS2 nanosheets for high efficient overall water splitting, Electrochim. Acta, 2018, 269, 55–61 CrossRef CAS .
  274. X. Q. Du, Z. Yang, Y. Li, Y. Q. Gong and M. Zhao, Controlled synthesis of Ni(OH)2/Ni3S2 hybrid nanosheet arrays as highly active and stable electrocatalysts for water splitting, J. Mater. Chem. A, 2018, 6, 6938–6946 RSC .
  275. Q. Xu, W. Gao, M. Wang, G. Yuan, X. Ren, R. Zhao, S. Zhao and Q. Wang, Electrodeposition of NiS/Ni2P nanoparticles embedded in amorphous Ni(OH)2 nanosheets as an efficient and durable dual-functional electrocatalyst for overall water splitting, Int. J. Hydrogen Energy, 2020, 45, 2546–2556 CrossRef CAS .
  276. L. M. Ren, C. Wang, W. Li, R. H. Dong, H. X. Sun, N. Liu and B. Y. Geng, Heterostructural NiFe-LDH@Ni3S2 nanosheet arrays as an efficient electrocatalyst for overall water splitting, Electrochim. Acta, 2019, 318, 42–50 CrossRef CAS .
  277. C. Hu, L. Zhang, Z. J. Zhao, A. Li, X. Chang and J. Gong, Synergism of Geometric Construction and Electronic Regulation: 3D Se-(NiCo)Sx/(OH)x Nanosheets for Highly Efficient Overall Water Splitting, Adv. Mater., 2018, 30, 1705538 CrossRef .
  278. B. Ma, X. Guo, X. Zhang, Y. Chen, X. Fan, Y. Li, F. Zhang, G. Zhang and W. Peng, Intercalated Graphite between Ni Foam and Ni3S2 Nanocrystals for the Activity Promotion in Overall Water Splitting, Energy Technol., 2019, 7, 1900063 CrossRef .
  279. H. C. Tsai, B. Vedhanarayanan and T. W. Lin, Freestanding and Hierarchically Structured Au-Dendrites/3D-Graphene Scaffold Supports Highly Active and Stable Ni3S2 Electrocatalyst toward Overall Water Splitting, ACS Appl. Energy Mater., 2019, 2, 3708–3716 CrossRef CAS .

This journal is © The Royal Society of Chemistry 2021