Interface engineering of transitional metal sulfide–MoS2 heterostructure composites as effective electrocatalysts for water-splitting

Yanqiang Li *a, Zehao Yin a, Ming Cui a, Xuan Liu a, Jiabin Xiong b, Siru Chen *b and Tingli Ma *cd
aState Key Laboratory of Fine Chemicals, School of Petroleum and Chemical Engineering, Dalian University of Technology, Panjin Campus, Panjin 124221, China. E-mail: yanqiangli@dlut.edu.cn
bCenter for Advanced Materials Research, Zhongyuan University of Technology, Zhengzhou, 450007, China. E-mail: siruchen@zut.edu.cn
cDepartment of Materials Science and Engineering, China Jiliang University, Hangzhou, 310018, China. E-mail: tinglima123@cjlu.edu.cn
dGraduate School of Life Science and Systems Engineering, Kyushu Institute of Technology, Kitakyushu, Fukuoka 808-0196, Japan. E-mail: tinglima@life.kyutech.ac.jp

Received 5th November 2020 , Accepted 26th November 2020

First published on 27th November 2020


Abstract

Benefiting from the high electrochemical surface area brought by the 2D nanosheet structure, MoS2 has received great research attention for the hydrogen evolution reaction (HER). Recently, it has been demonstrated that by constructing a transitional metal sulfide–MoS2 heterostructure, the HER performance of the MoS2-based catalysts can be further improved. It is even possible to obtain bifunctional catalysts for both HER and oxygen evolution reaction (OER) due to the synergistic effect of the different components in the composite, the electronic effect to enable an efficient electron transfer and appropriate binding energy for the intermediates of the electrocatalytic reactions, and the surface defects on the interface of the heterostructures. Herein, we review the recent progress on the construction of the transitional metal sulfide–MoS2 heterostructure for water splitting based on non-self-supporting and self-supporting catalysts. The surface and interface parameters of the heterostructures are discussed in detail to reveal the key roles of the hybrid structures for energy conversion. We also pay special attention to the theoretical simulations based on first principles to clarify the relationships between the electrochemical performance and structure parameters. Finally, the prospects and challenges of the transition metal sulfide–MoS2 heterostructures for water splitting in the future are proposed to prompt the reasonable design of transition metal sulfide–MoS2 heterostructures for full water splitting.


image file: d0ta10815d-p1.tif

Yanqiang Li

Dr Yanqiang Li received his PhD degree at Jilin University in 2015. Then, he joined Dalian University of Technology as a lecturer. His research focuses on the design and preparation of functional materials and catalysts for electrochemical energy storage and conversion based on the combination of theoretical studies and electrochemical experiments.

image file: d0ta10815d-p2.tif

Zehao Yin

Zehao Yin is a Master's degree candidate at Dalian University of Technology. He received her bachelor's degree from Xi'an University of Science and Technology. His research focuses on designing bifunctional electrocatalysts for high-performance zinc–air batteries.

image file: d0ta10815d-p3.tif

Siru Chen

Dr Siru Chen received her PhD degree at Jilin University in 2015. After three years research in Dalian Institute of Chemical Physics as a post-doctor, she joined Zhongyuan University of Technology as a lecturer in 2019. Her research focuses on the design and preparation of functional materials and catalysts for electrochemical energy storage and conversion.

image file: d0ta10815d-p4.tif

Tingli Ma

Professor Tingli Ma received her PhD degree in 1999 from the Department of Chemistry, Faculty of Science of Kyushu University, Japan. She then joined the National Institute of Advanced Industrial Science and Technology (AIST) as a Postdoctoral Researcher from 1999 to 2004. She worked as a professor at the Dalian University of Technology, China from 2007 to 2018. She is currently working at the Graduate School of Life Science and Systems Engineering at Kyushu Institute of Technology, Japan and China Jiliang University. She leads research teams studying inorganic and organic solar cells, such as dyesensitized solar cells and perovskite solar cells, and other related projects, including development of catalysts, hydrogen production, functional dyes and nano-semiconductor materials.


1. Introduction

The ever-increasing energy demand and depletion of fossil energy promote research studies to explore clean and renewable energy sources.1,2 Due to its carbon-free nature, hydrogen is considered an alternative clean energy carrier and has received widespread attention. However, there is no natural H2 on Earth, and the hydrogen element exists in the form of compounds. As one of the most important H-containing compounds, water is the most extensive resource on the earth. According to calculations, if all of the hydrogen in seawater is extracted, the total heat generated by hydrogen is thousands of times higher than the heat released by all fossil fuels on the earth.3–5 Therefore, the electrochemical splitting of water is a significant strategy to produce hydrogen. In addition, compared with the widely utilized methods to prepare hydrogen, such as steam reforming and coal gasification, the electrochemical electrolysis can produce purer hydrogen with no emission of greenhouse gases and other polluted gases.6,7

Electrochemical water splitting is a very mature technology, in theory, and two half reactions are involved: oxygen evolution reaction (OER) on the anode and hydrogen evolution reaction (HER) on the cathode.8 The thermodynamic voltage for water splitting is 1.23 V at 25 °C and 1 atm.9 However, in practical applications, a much larger potential is needed to drive this reaction due to the sluggish electrochemical kinetics. Nowadays, the energy conversion efficiency of water splitting is only 56–73%, and only less than 5% hydrogen is produced by this technology.10 Therefore, it is necessary to develop efficient catalysts to accelerate the reaction rate of HER and OER. So far, Pt-based catalysts are demonstrated to be the most efficient HER catalysts,11 while Ir and Ru oxides are regarded as the most efficient OER catalysts.12 However, due to the scarcity and high price of precious metals on the earth, their large-scale commercial application is limited.13 The development of catalysts with abundant earth reserves and excellent catalytic performance has become a critical factor in the commercialization of water splitting devices.

During the exploration process, transition metal-based compounds have been greatly investigated as electrocatalysts for efficient water splitting due to their earth-abundance and decent activity.14–17 Among them, MoS2 has received a lot of attention due to its high HER performance. MoS2 shows a typical layered structure similar to graphene, with weak van der Waals interactions between adjacent S–Mo–S layers, where Mo atoms are sandwiched between sulfur atoms in a hexagonal arrangement.18 The edge position of the MoS2 layered structure has been proved by theoretical calculations to be the active site for HER. Its edge also has a large number of unsaturated S atoms, which has an adsorption free energy close to 0 for the intermediate product in the HER process, making H protons easily adsorbed and desorbed on the surface of MoS2 to generate hydrogen.19,20 When used as a HER catalyst, only a small overpotential is required for MoS2 to achieve a higher current density. However, the surface layer of MoS2 is considered to be inert, which makes the process of how to expose more edge sites a critical issue. In addition, the intrinsic conductivity of MoS2 is relatively poor, which makes the resistance of the charge transport larger, resulting in larger potentials for the reaction.21 Moreover, the poor OER performance of MoS2 makes it unable to be used as a bifunctional catalyst to catalyze the water electrolysis.

At present, constructing metal sulfide–MoS2 heterostructures through interface engineering has been demonstrated as one effective strategy to improve the catalytic performance of MoS2-based catalysts.22 Due to the different catalytic activities toward HER and OER of the different components in the composites, it is probable to obtain bifunctional electrocatalysts for both OER and HER. In addition, heterostructures usually display increased electrocatalytic activities over the single-phase counterpart due to more exposed active sites on the interface, and the synergetic effect of the different components.23 Moreover, the interaction between the various components will cause the redistribution of electrons at the heterogeneous interface, so that it has a moderate adsorption-free energy for H proton and O-containing intermediate, which is very beneficial for boosting the HER and OER performance of the composite.24–27 Due to the rapid development of metal sulfide–MoS2 heterostructures for water splitting, a review about interface engineering of the MoS2-based electrocatalysts is necessary and meaningful to summarize the recent research progress on their design and application.

In this review, we focus on the construction of transition metal sulfide–MoS2 heterostructures as bifunctional electrocatalysts for overall water splitting. After the introduction on the background of the catalytic mechanism, the preparation process and structural characteristics of the materials are outlined based on non-self-supporting and self-supporting catalysts. The enhancement effects of the heterogeneous interfaces on the HER/OER performances of the catalysts are addressed, and the possible catalysis mechanism is discussed. Finally, the prospects and challenges of transition metal sulfide–MoS2 heterostructures for water splitting in the future are proposed. It is anticipated that this review will promote the reasonable design of transition metal sulfide–MoS2 heterostructures for full water splitting and other electrochemical energy conversion systems (Tables 1 and 2).

Table 1 Summary of non-self-supporting catalysts for HER, OER and water splitting
Electrocatalyst Electrolyte η @ j (mV @ mA cm−2) Tafel slope (mV dec−1) Overall voltage (V @ mA cm−2) Stability Ref.
HER OER HER OER
Co9S8@MoS2/CNFs 0.5 M H2SO4 for HER 190 @ 10 430 @ 10 110 61 Small changes in current density after 1000 CV cycles 47
1 M KOH for OER
Co9S8/NC@MoS2 0.5 M H2SO4 for HER 117 @ 10 68.8 12 h @ 67 mV in 1 M KOH 51
1.0 M PBS for HER 261 @ 10 126.1 12 h @ 117 mV in 0.5 M H2SO4
1 M KOH for HER 67 @ 10 60.3 12h @ 261 mV in 1.0 M PBS
MCS-HP 0.5 M H2SO4 for HER 175 @ 10 410 @ 10 in 1 M KOH 55.6 90.1 53
1 M KOH for HER 220 @ 10 115.3
Co9S8/MoS2@NSOC 0.5 M H2SO4 for HER 233 @ 10 96 With the loss of 11% and 9.4% respectively after 12 h work in the acidic and alkaline media 55
1 M KOH for HER 194 @ 10 118
CoS2–C@MoS2-25 0.5 M H2SO4 for HER 173 @ 10 391 @ 10 61 46 Current density remained virtually immobile for 1000 cycles 56
1 M KOH for OER
CoS1.097/MoS2 0.5 M H2SO4 for HER 228 @ 10 59 21 h @ 10 mA cm−2 in acidic, neutral and alkaline media for HER 57
0.1 M PBS for HER 341 @ 10 85
1 M KOH for HER 249 @ 10 75
Ni9S8@MoS2 0.5 M H2SO4 for HER 88 @ 10 49 20 h @ 10 mA cm−2 for HER 64
NiS2/MoS2-RGO 0.5 M H2SO4 for HER 172 @ 10 51 10 h @ 10 mA cm−2 in 0.5 M H2SO4 and 1 M KOH for HER 68
1.0 M PBS for HER 229 @ 10 103
1 M KOH for HER 144 @ 10 82
MoS2–NiS2/NGF 1.0 M KOH 172 @ 10 370 @ 10 70 1.64 @ 10 24 h @ 10 mA cm−2 73
CuS@defect-rich MoS2 0.5 M H2SO4 for HER 135 @ 10 50 10[thin space (1/6-em)]000 s @ 175 mV for HER 76
ZnS@C@MoS2 1 M KOH for HER 118 @ 10 55.4 12 h @ 30 mA cm−2 for HER 80
MoS2 NDs/VS2 0.5 M H2SO4 for HER 291 @ 10 58.1 16 h @ 300 mV for HER 83
NiCo2S4@MoS2 1 M KOH for HER 194 @ 10 62 10 h @ 10 mA cm−2 for HER 87
MoS2/NiCoS 1 M KOH 189 @ 10 290 @ 10 75 77 1.50 @ 10 22 h @ 10 mA cm−2 88
Co,Fe–MoS2 1 M KOH 113 @ 10 264 @ 10 61 117 1.49 @ 10 10 h @ 10 mA cm−2 89
MoS2–g-CuNi2S4 0.5 M H2SO4 for HER 120 @ 10 29.3 Small changes in current density after 1000 CV cycles 90


Table 2 Summary of self-supporting catalysts for HER, OER and water splitting
Electrocatalyst Electrolyte Substrate η @ j (mV @ mA cm−2) Tafel slope (mV dec−1) Overall voltage (V @ mA cm−2) Stability Ref.
HER OER HER OER
Co5Mo10Sx/CC 1 M KOH Carbon fiber cloth 36 @ 10 153 @ 10 56 58 1.51 @ 10 12 h @ 10 mA cm−2 105
MoS2/CoS2 0.5 M H2SO4 for HER Carbon fiber paper 126 @ 10 42.6 10 h @ 10 mA cm−2 106
1.0 M PBS for HER 196 @ 10 77.5
1 M KOH for HER 97 @ 10 78.7
CoS2/MoS2 HNSAs 1 M KOH for HER Ni foam 50 @ 10 76 70 h @ 100 mV 110
CoMoSx/NF 1 M KOH Ni foam 89 @ 10 345 @ 100 94 1.89 @ 500 25 h @ 500 mA cm−2 113
Mo-doped Ni3S2 1 M KOH Ni foam 212 @ 10 260 @ 10 98 85 1.67 @ 10 48[thin space (1/6-em)]000 s @ 10 mA cm−2 119
MoS2–Ni3S2 HNRs/NF 1 M KOH Ni foam 98 @ 10 290 @ 10 61 57 1.50 @ 10 48 h @ 17 mA cm−2 120
MoS2/Ni3S2 heterostructures 1 M KOH Ni foam 110 @ 10 218 @ 10 83 88 1.56 @ 10 10 h @ 10 mA cm−2 121
N–NiS/MoS2 1 M KOH Ni foam 71 @ 10 79 20 h @ 30 mA cm−2 125
p-MoS2/Ni3S2/NF 1 M KOH Ni foam 99 @ 10 185 @ 10 71 46 35 h @ 10 mA cm−2 128
MoS2/FNS/FeNi foam 1 M KOH FeNi foam 120 @ 10 204 @ 10 45.1 28.1 10 h @ 10 mA cm−2 for HER 134
10 h @ 204 mV for OER
(Ni, Fe)S2@MoS2 1 M KOH Carbon fiber paper 130 @ 10 270 @ 10 101.2 43.2 1.56 @ 10 40 h @ 10 mA cm−2 135
CoMoNiS–NF-31 1 M KOH Ni foam 113 @ 10 166 @ 10 85 58 1.54 @ 10 24 h @ 20 mA cm−2 in 1 M KOH 138
0.5 M H2SO4 103 @ 10 255 @ 10 55 78 1.45 @ 10 Decreased from 21.7 to 17.4 mA cm−2 after 80 min in 0.5 M H2SO4
1.0 M PBS 117 @ 10 405 @ 10 56 71 1.80 @ 10 20 h @ 1.80 V in 1.0 M PBS
Mo(1–x)WxS2@Ni3S2 1 M KOH Ni foam 98 @ 10 285 @ 10 92 90 1.62 @ 10 50 h @ 10 mA cm−2 139


2. Mechanism for HER and OER

The HER reaction takes place at the cathode of the water electrolysis device, and mainly involves three reaction steps with two possible reaction mechanisms. One is the Volmer–Heyrovsky mechanism which contains the Volmer reaction and Heyrovsky reaction, and the other one is the Volmer–Tafel mechanism, which contains the Volmer reaction and Tafel reaction.28 In both reaction mechanisms, the first step is the adsorption of H on the active sites (acidic medium) or the decomposition of a water molecule (alkaline medium) and adsorption of H to form adsorbed H, which is called the Volmer reaction or the discharge reaction. The second step is either the adsorbed H combines one proton/H2O and an electron to form H2 (Heyrovsky reaction), or two adsorbed H atoms combine to generate H2 (Tafel reaction).29

Volmer reaction:

H3O+ + M + e → M–H + H2O (in acidic media)

H2O + M + e → M–H + OH (in alkaline media)

Heyrovsky reaction:

M–H + H3O+ + e → M + H2 + H2O (in acidic media)

M–H + H2O + e → H2 + M + OH (in alkaline media)

Tafel reaction:

2M–H → 2M + H2 (Tafel reaction in both acidic and alkaline media)

Regardless of the different reaction mechanisms, the adsorbed H is always involved in the reactions, and the free energy of H adsorption is an important parameter to evaluate the intrinsic activity of an electrocatalyst for HER. When the adsorption of H on the surface of the catalyst is too weak, the Volmer reaction is the rate-determination step. Conversely, when the adsorption of H on the surface of the catalyst is too strong, the desorption of H2 is the rate-determination step. Ideally, the free energy of H adsorption should be close to zero, like Pt.30

Compared with HER, OER (which occurs at the anode of the water electrolysis device) is a more complicated four-electron transfer process, and different intermediates (M–O, M–OH, M–OOH) are involved. The generally accepted reaction mechanisms in alkaline or acidic media are listed in the following equations. First, hydroxyl or water molecules are adsorbed on the catalyst to form M–OH. Then, M–OH continues to undergo an electron transfer to form M–O. After the formation of M–O, O2 can be produced via two routes. One is the direct generation of molecular oxygen from two M–O species, and the other involves M–O being oxidized into the M–OOH species first, then O2 is released from the M–OOH intermediate.31–34

In acidic media:

M + H2O → M−OH + H+ + e

M−OH → M−O + H+ + e

2M−O → 2M + O2

M−O + H2O → M−OOH + H+ + e

M−OOH → M + O2 + H+ + e

In alkaline media:

M + OH → M−OH + e

M−OH + OH → M−O + H2O + e

2M−O → 2M + O2

M−O + OH → M−OOH + e

M−OOH + OH → M + O2 + H2O + e

Like HER, OER involves the adsorption and desorption of oxygen-containing intermediates on the active sites.35,36 When the adsorption of O on the surface of the catalyst is too weak, the M–OH species is not easily formed. In contrast, when the adsorption of O on the active sites is too strong, M–OOH and O2 are not easy to generate. Therefore, the optimum of the adsorption free energy for the O-containing intermediates can be anticipated to enhance the OER performance of the catalysts.37–39

3. Heterostructure catalysts

3.1 Chemical synthetic strategy of transitional metal sulfide–MoS2

The chemical synthetic strategies of transitional metal sulfide–MoS2 heterostructure catalysts mainly fall into two types. One is the direct synthesis, where the metal sources and S sources are dissolved in a solvent, followed by the solvothermal reaction. Based on whether the supports are added, self-supporting or nonself-supporting metal sulfide–MoS2 heterostructures can be facilely synthesized. For example, using cobalt acetate and sodium molybdate as the Co and Mo sources, followed by the solvothermal reaction in the presence of thiourea, Cao and co-workers prepared Co9S8@MoS2 hybrid nanospheres for overall water splitting.40 Liu and co-workers prepared a self-supported MoS2/Ni3S2 heterostructure catalyst by a facile one-step hydrothermal method, where sodium molybdate, Ni foam and thioacetamide were used as raw materials.26

The other type of synthesis strategy is a two-step reaction route, where bimetallic precursors are prepared first, and the sulfurization reaction was carried out subsequently to prepare the metal sulfides. The bimetallic precursors can be metal oxides, metal hydroxides or metal organic frameworks, and the sulfurization reaction can be a hydrothermal reaction or thermal annealing in the presence of sulfur sources. For example, using ZIF-67 as the precursor, Sun and co-workers prepared carbon supported Co9S8@MoS2 hybrids by thermal annealing and sequenced solvothermal reaction.41 By in situ growth of Cu(OH)2 nanorod arrays on Co foam, and further hydrothermal reaction with sodium molybdate and thiourea, Yu and Co-workers prepared a Cu foam supported Cu2S/MoS2 heterostructure for HER.42

3.2 Features and advantages of heterostructure catalysts

Compared with single component catalysts, heterostructure catalysts that are composed of multicomponent have unique advantages. First, the synergistic effect occurred when a catalyst contains a multicomponent, and this can be fully utilized to develop bifunctional electrocatalysts by introducing H2O/OH adsorption sites. A well-established viewpoint is that MoS2 is very efficient for HER, and Co, Ni and Fe-based sulfides are beneficial for OER. Therefore, by forming heterostructure composites, the catalysts can be used for overall water splitting.

Second, the electronic effect is another factor that boosts the catalytic activity of heterostructure catalysts. This is because the strong electronic interaction between two different active components will efficiently reconstruct the active centers, which is favourable for yielding high catalytic activity. For example, Shao and co-workers observed electron transfer from (Ni, Fe)S2 to MoS2 in (Ni, Fe)S2@MoS2 heterostructures by X-ray photoelectron spectroscopy, and it is suggested that the electronic interaction between (Ni, Fe)S2 and MoS2 could weaken S–Hads formed on the interface, which regulates H chemisorption to facilitate the Volmer–Heyrovsky process.43 In other works, the roles of the electronic coupling effect were also observed, such as lowering the energy barrier of intermediates or achieving an appropriate binding energy for the electrocatalytic reactions.

Third, the defect effects are also observed in heterostructure catalysts. Introducing defects into the electrocatalysts is also an effective method to improve their catalytic performance. In heterostructure catalysts, the long-range orderly structures of the active species are interrupted due to the mismatch lattice arrangement; therefore, many defects can be formed. For example, Du and coworkers prepared a Co9S8/Ni3S2 heterostructure with a large number of lattice defects in the interface, and density functional theory indicates that the abundant lattice defects could boost the chemisorption of the intermediates, which greatly promoted its overall water splitting performance.44

3.3 Characterization of heterostructure catalysts

Accurately characterizing the interface interaction of heterostructure catalysts is crucial for designing and developing more efficient materials for catalysis. Generally, the various metal sulfide species and their crystal phases can be inferred from X-ray diffraction and transmission electron microscopy (TEM) techniques. In addition, high-resolution TEM images can be used to observe the intimate contact and the formed interfaces between different objects or phases. X-ray photoelectron spectroscopy (XPS) and ultraviolet photoemission spectroscopy can be utilized to characterize the electron interaction between the different species, and electron paramagnetic resonance (EPR) can be used to investigate the defects in the heterostructure catalysts. Combined with a more accurate theoretical model, the structures and interface interaction of heterostructure catalysts can be well established (Fig. 1 and 2).
image file: d0ta10815d-f1.tif
Fig. 1 A summary of various transition metal sulfide–MoS2 heterostructure composites.

image file: d0ta10815d-f2.tif
Fig. 2 Schematic diagram of the reaction mechanism for water splitting at the cathode (HER) and anode (OER).

4. Non self-supporting metal sulfide–MoS2 heterostructure composites

4.1 CoSx–MoS2 heterostructure composites

Cobalt sulfides, including Co9S8, Co3S4, CoS2 and CoS, have been demonstrated as attractive electrocatalysts for water splitting owing to their high catalytic activity, as well as excellent electrochemical stability for HER and OER.45 When cobalt sulfides and MoS2 are combined together to form reasonable and orderly heterostructures, the synergistic effect between the two components can enhance the overall electrochemical catalytic performance, which is better than the performance of each single component.46 In 2015, using polyacrylonitrile (PAN) nanofibers with Co(NO3)2 and ammonium tetrathiomolybdate (NH4)2MoS4 as the precursor, Zhu et al. synthesized a cubic cobalt sulfide-layered molybdenum disulfide core–shell/carbon nanofiber (Co9S8@MoS2/CNFs) composite.47 The Co9S8@MoS2 core–shell nanostructure is composed of a Co9S8 nanoparticle core and fullerene-like MoS2 shell. As shown in Fig. 3a and b, the authors investigated the HER pathways on the surface of MoS2 and Co9S8@MoS2 through first-principles calculations. Although the Co9S8@MoS2 catalyst exhibits almost the same H atom adsorption energy compared with MoS2, the chemical binding of the hydrogen molecules on the MoS2 surface is much larger. Therefore, the desorption of H2 from Co9S8@MoS2 is easier, proving that by constructing the Co9S8/MoS2 hetero-interface, improved catalytic activity can be achieved due to the reduced energy barrier for the desorption of the H2 molecule.
image file: d0ta10815d-f3.tif
Fig. 3 (a) The kinetic energy barrier profiles of HER on the MoS2 (top) and Co9S8@MoS2 (down) catalysts. The right side shows the diffusion paths of H2 on the MoS2 and Co9S8@MoS2 surfaces. (b) Optimized structures for the top view of H–MoS2. Reproduced with permission.47 Copyright 2015, Wiley-VCH. (c) Schematic illustration for the synthesis of Co9S8/NC@MoS2 polyhedrons. (d) LSV curves of NC, Co/NC, MoS2, Co9S8/NC, Co9S8/NC@MoS2, and Pt/C in 1.0 M KOH and 0.5 M H2SO4. (e) Polarization curves of the Co9S8/NC@MoS2 before and after 2000 CV cycles in 1.0 M KOH solution. Reproduced with permission.51 Copyright 2017, American Chemistry Society. (f) Schematic illustration of the fabrication of MCS-HP. (g) Mo 3d and S spectra in MoS2 and MCS-HP. (h) Co 2p in CS-HP, and MCS-HP. Reproduced with permission.53 Copyright 2018, Elsevier.

Recently, metal–organic frameworks (MOFs), which consist of metal ion nodes and organic linkers, have attracted increased attention due to their well-defined pore structure, high surface area, and tunable porosity.48 Due to the weak coordination bonding and regular morphology of MOFs, they have been widely utilized as good precursors for the synthesis of nanostructured electrode/catalyst materials.49 As a well-known Co-based MOF, ZIF-67 has been widely investigated as a precursor for preparing Co-based catalysts.50 Li et al. synthesized hierarchical porous Co9S8/N-doped carbon@MoS2 (Co9S8/NC@MoS2) polyhedrons for HER over a broad pH range by using ZIF-67 and ammonium molybdate as precursors (Fig. 3c).51 Co9S8/NC@MoS2 retains the polyhedral structure of ZIF-67, and the layered MoS2 nanosheets are uniformly coated on the Co9S8/NC polyhedron. The Co/NC polyhedron derived by the carbonization of ZIF-67 has high porosity and active surface, which is beneficial for the nucleation and growth of MoS2 nanosheets on the Co/NC polyhedron. The coupling interface between MoS2 and the conductive carbon framework can reduce the work function of the carbon layers and optimize the electron structure, which is beneficial to accelerate the transfer of electrons. In addition, combining Co9S8 and MoS2 into a reasonably designed structure can combine the inherent properties of these two materials together, enabling Co9S8/NC@MoS2 with excellent HER activity over a broad pH range, with overpotentials of 67 at 10 mA cm−2 in 1.0 M KOH (Fig. 3d). In addition, the Co9S8/NC@MoS2 electrode exhibits an overpotential increase of only 3 mV after 2000 CV cycles to afford a current density of 10 mA cm−2, suggesting its good stability (Fig. 3e).

In addition to Co9S8/MoS2 composite materials, ZIF-67 has been used as a precursor to prepare Co3S4/MoS2 materials.52 For example, using ZIF-67 and sodium molybdate as raw materials, Lei et al. synthesized a MoS2/Co3S4 hollow polyhedral (MCS-HP) co-catalyst through the templated solvothermal method (Fig. 3f).53 The synthesized MCS-HP exhibits a well-defined polyhedral structure with molybdenum sulfide nanosheets covering the surface of Co3S4. Through XPS characterization, it can be observed that the Mo and Co peaks have shifted. The MoS2 that received electrons is in an electron-rich state, and the binding energy of Mo 3d is negatively shifted due to the shielding effect. On the contrary, due to the penetration effect, the Co 2p binding energy in the electron-deficient Co3S4 increased, demonstrating the strong interaction between MoS2 and Co3S4 (Fig. 3g and h). By appropriately modifying the d orbital electrons of the transition metal sulfides, the electrochemical catalytic activity of the catalyst can be improved. It only needs overpotentials of 175 and 220 mV to achieve a current density of 10 mA cm−2 in acid and alkali media, respectively.

In previous work, the MOF-templated synthesis of cobalt sulfide-based materials is limited due to the choice of ZIF-67 as the precursor, which requires an additional sulfurization process.54 Recently, Chen prepared the bimetallic sulfide Co9S8/MoS2 uniformly supported on N,S,O co-doped carbon (Co9S8/MoS2@NSOC) through the one-step pyrolysis of the Co-based MOF (Co-NOMOF) with thiomolybdate [Mo3S13]2− (Fig. 4a).55 The three-dimensional structure of Co-NOMOF facilitates the entry of Co nanoparticles into the carbon layer, and promotes the in situ doping of N and O atoms. Conversely, [Mo3S13]2− can introduce Mo and S sources, and further result in S doping into the carbon matrix. At the same time, it promotes the occurrence of the vulcanization process, and a Co9S8/MoS2 heterogeneous was finally achieved. The porosity of the Co9S8/MoS2@NSOC materials was investigated by N2 adsorption at 77 K, and the results show that there are mesopores and micropores in Co9S8/MoS2@NSOC. In addition, the calcination temperatures will significantly affect the pore structure of MOF-derived products. A high carbonization temperature is more beneficial to the formation of a porous structure. However, an excessively high temperature (1000 °C) will damage the porous structure, especially the mesopores to a certain extent (Fig. 4b and c). The rich pore structure can bring a large specific surface area to facilitate the transfer of electrolyte, and provide more active sites. Co9S8/MoS2@NSOC-900 exhibited remarkable catalytic performance for HER with overpotentials of 194 and 233 mV in 1 M KOH and 0.5 M H2SO4 solution at 10 mA cm−2, respectively, which is much better than the comparable catalysts, such as Co9S8, MoS2 and Co9S8/MoS2 (Fig. 4d and e). Its excellent performance can be ascribed to the unique electronic structure and lower local work function of the multi-heteroatom-doped carbon nanomaterials, the synergy between Co9S8 and MoS2, and the Co9S8/MoS2 heterogeneous interface that promotes the dissociation step of H2O.


image file: d0ta10815d-f4.tif
Fig. 4 (a) Schematic illustration for the synthesis of Co9S8/MoS2@NSOC. (b) N2 adsorption–desorption isotherms of Co9S8/MoS2@NSOC. (c) Pore size distribution curves of the Co9S8/MoS2@NSOC catalysts. (d and e) LSV curves of the Co9S8/MoS2@NSOC900, Co@PC, MoS2, Co9S8, Co9S8–MoS2 materials for HER in 0.5 M H2SO4 and 1.0 M KOH solution. Reproduced with permission.55 Copyright 2018, Royal Society of Chemistry. (f) Schematic illustration of the synthesis of CoS2–C@MoS2. Reproduced with permission.56 Copyright 2019, American Chemistry Society. (g) Schematic illustration for the synthesis of CoS1.097/MoS2. Reproduced with permission.57 Copyright 2019, American Chemistry Society.

In addition to the MOF-templated synthesis strategy, other precursors were selected to synthesize cobalt–molybdenum bimetallic sulfides with regular morphologies. Through the one-step hydrothermal and vulcanization treatment of Co–C nanofibers, Zhu et al. synthesized hierarchical CoS2–C@MoS2 core–shell nanofibers (Fig. 4f).56 The electrochemical measurements demonstrated that the conductivity and exposure of the active centers are two key factors that affect the catalytic performance of the catalyst. CoS2–C makes up for the shortcomings of MoS2 on conductivity, and has a synergistic effect with MoS2 to improve the performance of the catalyst. Sun et al. used cobalt chloride and ammonium molybdate as raw materials to synthesize porous CoS1.097/MoS2 hybrid microspheres with rich defects through a one-step vulcanization (Fig. 4g).57 Due to the synergy between CoS1.097 and MoS2, abundant defects and hierarchical interconnected microporous/mesoporous structure, the electrocatalytic activity of the CoS1.097/MoS2 hybrid microspheres show good HER performance, with overpotentials of 228, 249, and 341 mV at 10 mA cm−2 current density in 0.5 M H2SO4, 1 M KOH, and 0.1 M PBS electrolyte, respectively.

As shown above, for the synthesis of the CoSx–MoS2 heterostructure composites, direct hydrothermal reaction and MOFs templated method are usually utilized. For the direct hydrothermal reaction method, the obtained metal sulfides are usually sphere particles, which greatly restrict the exposure of active sites. Although hollow metal sulfides can be readily prepared by the MOFs templated strategy, it should be noted that the low yield and high cost of MOFs would prohibit the large-scale preparation of the composites. Therefore, a more effective synthetic strategy should be developed for the commercial application of the composites.

4.2 NiSx–MoS2 heterostructure composites

Nickel sulfides are considered to be highly efficient water electrolysis catalysts due to their excellent electrical conductivity.58–60 However, the low density of the electrocatalytic sites on the surface and the low stability inhibit its further application.61 Combining nickel sulfide with MoS2 can produce a synergistic effect by constructing a heterogeneous interface and enrich the active sites, and great progress was achieved in this area.62,63

Using MoS2 for the in situ growth of Ni9S8, Xu et al. prepared a Ni9S8@MoS2 composite for HER.64 Due to the electronic coupling between Ni9S8 and MoS2, the charge transfer resistance for Ni9S8@MoS2 is greatly reduced. The electrochemical surface area of Ni9S8@MoS2 is also much larger than that for Ni9S8 and MoS2, demonstrating the fast electron transfer and more exposed active sites of the Ni9S8@MoS2 heterostructure (Fig. 5a–c). As a result, Ni9S8@MoS2 exhibited good HER catalytic activity, with an overpotential of 88 mV at 10 mA cm−2 and a low Tafel slope of 49 mV dec−1.


image file: d0ta10815d-f5.tif
Fig. 5 (a) Schematic illustration of the heterostructure Ni9S8@MoS2 hybrid. (b) EIS Nyquist plots of Ni9S8@MoS2 and other materials. (c) Estimation of Cdl by plotting the current density variation vs. scan rate to fit a linear regression. Reproduced with permission.64 Copyright 2019, Elsevier. (d) Schematic illustration for the formation of NGF and MoS2–NiS2/NGF. (e and f) HRTEM images of MoS2–NiS2/NGF. (g) Polarization curves of MoS2–NiS2/NGF, MoS2–NiS2, NGF, Pt/C, and RuO2 in 1.0 M KOH for HER and OER. (h) Polarization curves of two-electrode overall water splitting using MoS2–NiS2/NGF/NF as both anode and cathode in 1.0 M KOH. Inset: galvanostatic measurement of the MoS2–NiS2/NGF/NF performed at a current density of 10 mA cm−2 in 1.0 M KOH. Reproduced with permission.73 Copyright 2019, Elsevier.

The catalytic efficiency of the catalyst is affected by its conductivity. Poor conductivity will cause additional overpotentials, which will consume more energy and reduce the apparent activity of the catalyst.65–67 The combination of the MoS2/NiSx heterogeneous catalysts with high conductive carbon materials is expected to further improve the performance of the catalysts. Wang et al. synthesized a NiS2/MoS2 composite on reduced graphene oxide using NiMoO4–RGO as the precursor,68 and the tree-like NiS2/MoS2–RGO nanocomposites afford a current density of 10 mA cm−2 at small overpotentials of 172, 144 and 229 mV in acidic, alkaline and neutral solution, respectively. Xu et al. first treated graphene oxide with NH3 to obtain the N-doped reduced graphene oxide, and then loaded Ni3S4/MoS2 on the carbon support.69 The optimized sample possesses a large number of catalytic active sites for HER and high conductive 3D networks that benefit electron transfer, and excellent catalytic activity was achieved.

Integrating 2D carbon materials with metal sulfides can improve the conductivity of the materials. However, due to the van der Waals interaction and strong π–π accumulation between the graphene layers, the 2D graphene is easy to aggregate. This irreversible aggregation not only hinders the uniform dispersion of the catalyst, but also makes the surface inaccessible to the reactants and electrolytes, which is not beneficial for the improvement of the catalytic performance.70 In comparison, 3D graphene is an ideal scaffold for improving the electrocatalytic activity of the supported materials due to its high macroporosity and high conductivity.71,72 Using melamine foam for the adsorption of graphene oxide and further burning off the melamine template, Kuang et al. first prepared a 3D interconnected macroscopic porous nitrogen-doped graphene foam (NGF).73 The surface of the graphene foam with a negative charge can electrostatically adsorb Mo2+ and Ni2+ cations to form MoO2–Ni(OH)2 nanowalls on the NGF framework. Finally, after a CVD process, bimetallic sulfides of MoS2–NiS2 anchored on the 3D nitrogen-doped graphene foam (NGF) was successfully prepared and used as a bifunctional catalyst for water splitting (Fig. 5d). In comparison, the MoS2–NiS2 prepared without NGF exhibited severe agglomeration, which significantly reduced the contact area between the electrolyte and the catalyst, and hindered the release of gas products. The 3D hollow tube structure of MoS2–NiS2/NGF ensures abundant active sites and diverse paths to achieve the rapid and efficient transport of the electrolyte and gas. The TEM images clearly show that the heterointerface is formed between the (002) plane of MoS2 and (210) plane of NiS2 (Fig. 5e and f). In the electrolyte, H protons tend to be adsorbed by MoS2, while OH tends to be adsorbed by NiS2, resulting in the rapid dissociation step of H–O at the heterogeneous interface. The special catalyst delivered a current density of 10 mA cm−2 at overpotentials of 172 mV for HER and 370 mV for OER in 1.0 M KOH. Furthermore, when used in an alkaline water electrolyzer, MoS2–NiS2/NGF can produce the vigorous and continuous evolution of H2 and O2 at a current density of 10 mA cm−2 under a cell voltage of 1.64 V (Fig. 5g and h).

Although integrating carbon materials with metal sulfides could effectively improve the conductivity of the catalysts and prohibit the aggregation of MoS2, it should be noted that the real charming fascinating of MoS2 is its 2D nanosheet structure, which can expose the maximum number of active sites. Although it is possible to obtain several layers of MoS2 by exfoliation technology, the preparation of the ultrathin MSx–MoS2 nanosheet is rarely reported. It is also very significant to synthesize the few-layers and even the single layer MSx–MoS2 heterostructure catalysts to boost their catalytic performances.

4.3 MSx–MoS2 (M = Cu, Zn, V) heterostructure composites

Due to its anisotropic properties, CuS can provide a penetration channel for ion adsorption and transmission.74 Meanwhile, the S atom in CuS contains empty 3p orbitals and a large number of electrons, which has a strong tendency to capture electrons and promote electron transfer reactions with electron donor molecules.75 Therefore, designing CuS and MoS2 composites could facilitate electron transfer during the catalytic process. In addition, introducing defects into catalysts is a valid strategy to promote their catalytic activity. In a recent study, Liu et al. synthesized a CuS@defect-rich MoS2 (CuS@MoS2-DR) core–shell structure and investigated its catalytic activity for HER.76 The CuS nanosphere was synthesized first, and SiO2 was coated on the surface of CuS for the further growth of MoS2 by using sodium molybdate and thiourea as raw materials. After removing SiO2 by HF etching, CuS@MoS2-DR was prepared. By changing the molar ratio of sodium molybdate and thiourea, the authors found that excess thiourea can be adsorbed on the surface of MoS2 to hinder the growth of oriented crystals, thereby forming a defect structure with many dislocations, distortions and the cracking of the basal planes (Fig. 6a). In the HER catalysis process, CuS can effectively trap the excess electrons of defect-rich MoS2, and then quickly transport them to the core–shell structure (Fig. 6b). Through EIS characterization, it can be found that the integration of CuS and MoS2 can significantly reduce the charge transport resistance, and CuS@MoS2-DR-0.4 with 40% MoS2 loading has the best conductivity (Fig. 6c). Benefiting from the additional active edge sites of the defect-rich MoS2 and the improved charge transport facilitated by the intimate interface, the CuS@defect-rich MoS2 could afford a current density of 10 mA cm−2 at a low overpotential of 135 mV (Fig. 6d).
image file: d0ta10815d-f6.tif
Fig. 6 (a) Schematic illustration for the synthesis of CuS@MoS2. (b) Schematic illustration of the charge transfer in CuS@MoS2-DR and the proposed mechanism for HER. (c) Nyquist plots of CuS@MoS2-DR with other materials. (d) Polarization curves of CuS@MoS2-DR with other materials. Reproduced with permission.76 Copyright 2020, Elsevier.

In addition to copper sulfide, transition metals ZnS have received extensive attention because of their excellent electrical conductivity for electrocatalysis.77–79 Liu et al. synthesized hierarchical ZnS@C@MoS2 core–shell nanostructures by the bottom-up two-step hydrothermal strategy combined selective etching method (Fig. 7a).80 As the core, ZnS is covered with a defective porous carbon shell, and ultra-thin MoS2 nanosheets are grown on the surface of the carbon shell. The coexistence of the ZnS semiconductor and the porous defect-rich carbon shell not only increases the conductivity of the material (Fig. 7b), but also separates and supports exfoliated MoS2 nanosheets, which can effectively expose a large number of catalytically active edge sites. The hierarchical ZnS@C@MoS2 core–shell nanostructures exhibited a low overpotential of 118 mV at 10 mA cm−2 for HER (Fig. 7c).


image file: d0ta10815d-f7.tif
Fig. 7 (a) Schematic representation for the synthesis of ZnS@C@MoS2. (b) Nyquist plots of ZnS@C@MoS2 with other materials. (c) Polarization curves of ZnS@C@MoS2 with other materials for HER in 1.0 M KOH. Reproduced with permission.80 Copyright 2019, Elsevier. (d) Schematic illustration of the synthesis of MoS2 NDs/VS2 and pristine VS2. Reproduced with permission.83 Copyright 2018, American Chemistry Society.

In addition, VS2 has gradually become a kind of promising electrocatalyst due to its unique electronic structure. Unlike other transition metal sulfides, VS2 shows an inherent metallic electronic ground state and unique electrical conductivity, making it a potential active HER electrocatalyst.81,82 Du et al. synthesized a layered heterostructure composed of MoS2 nanodots and metal VS2 nanosheets by in situ hydrothermal using water-soluble monolayer MoS2 nanodots as the size-controllable precursor, and sodium orthovanadate and thioacetamide as the sources of V and S, respectively.83 During the hydrothermal process, MoS2 nanodots act as a separator to inhibit the restacking of VS2 nanodots, while layered VS2 nanosheets act as a supporting skeleton to effectively hinder the agglomeration of MoS2 nanodots (Fig. 7d). MoS2 and VS2 can form a highly lattice-matched electron transfer heterojunction due to the orbital interaction. The redistribution of the charge is beneficial to stabilizing VS2, making it exhibit metallic characteristics and improve the overall conductivity. The optimized hybrid material exhibits significant electrocatalytic performance in acidic media due to its thinner flake heterostructure, higher conductivity, more catalytic active centers and higher intrinsic activity. Moreover, the LSV polarization curves before and after 1000 CV are almost the same, and the current density remains stable after 16 hours of chronoamperometry, demonstrating its good stability.

4.4 Non-self-supporting MoS2-based trimetallic sulfide composites

Constructing a ternary metal sulfides–MoS2 composite can further improve the catalytic activity of the multiple transition metal sulfides due to the more abundant interfaces and active sites.84 The heterogeneous interface in multiple transition metal sulfides can increase the electron density at the active edge, and the electronic structure of the catalysts can be further optimized due to the synergistic effect of various metal sulfides.85 Hu et al. designed MoS2/CoNi2S4 composites based on the theory-calculation guidance.86 By simultaneously modulating the 3d band offset of Ni, Co, and Mo near the interface, the author optimized the electronic structure of MoS2/CoNi2S4. According to dual-pathway kinetic analyses, it can be found that under acid–base conditions, the activation energy of MoS2/CoNi2S4 in both Heyrovsky and Volmer steps is slightly lower than that of the MoS2 nanosheet (Fig. 8a and b), indicating that MoS2/CoNi2S4 has better kinetics in acid and alkali media than MoS2. The hybridization of MoS2 and CoNi2S4 accelerates the dissociation step of water molecules induced by the charge transfer in an alkaline environment. DFT analysis showed that the electronic modulator Mo in MoS2/CoNi2S4 simultaneously modified Ni and Co atoms, which not only activated Ni, but also alleviated the over-binding effect from Co (Fig. 8c–e). Due to the well-defined electronic structure of the MoS2/CoNi2S4 composite, it could effectively adsorb H and OH in an alkaline environment to boost its overall water splitting performance.
image file: d0ta10815d-f8.tif
Fig. 8 (a and b) Comparison of the free energy diagram for the dominated Volmer–Heyrovsky pathways of MoS2 and MoS2/CoNi2S4 catalysts for HER in 0.5 M H2SO4 and 1 M KOH electrolytes, respectively. (c) The site-dependent PDOSs of Mo from the interface regions toward the metal Mo. (d) The site-dependent PDOSs of Ni from the interface toward metal Ni. CNS represents CoNi2S4 bulk. (e) The site-dependent PDOSs of Co from the IF regions toward the metal Co. Reproduced with permission.86 Copyright 2019, Wiley-VCH. (f) Schematic illustration of the synthesis of NiCo2S4@MoS2. (g and h) Polarization LSV curves and Tafel plots of NiCo2S4@MoS2 and NiCo2S4 in 1 M KOH. Reproduced with permission.87 Copyright 2019, Royal Society of Chemistry. (i) Schematic of the preparation of MoS2/NiCoS heterostructures. (j) Schematic illustration of the water splitting process for MoS2/NiCoS heterostructures. (k) Polarization curves for water splitting using MoS2/NiCoS and Ni foam as bifunctional catalysts in a 1.0 M KOH solution. Reproduced with permission.88 Copyright 2019, Royal Society of Chemistry.

Using ZIF-67 as template, Song et al. synthesized hollow core–shell NiCo2S4@MoS2 polyhedrons, where MoS2 nanosheets were assembled on NiCo2S4 nanobuilding blocks (Fig. 8f).87 Compared with NiCo2S4 dodecahedrons, NiCo2S4@MoS2 exhibits better HER performance under alkaline conditions (Fig. 8g and h), indicating that the synergy between NiCo2S4 and MoS2 is beneficial to increase the catalytic activity. Together with a 3D architecture of the outer MoS2, the hollow structure prevents the adhesion of H2 bubbles to the electrode surface, thereby keeping the integrity of the necessary solid–liquid interface. Qin et al. also synthesized MoS2/NiCoS nanosheets and investigated their catalytic performance for water splitting through a MOF-templated strategy.88 The authors synthesized NiCo–MOF nanosheets first, which then reacted with (NH4)2MoS4 to generate two-dimensional porous MoS2/NiCoS (Fig. 8i). The two-dimensional porous structure of MoS2/NiCoS can expose more unsaturated S boundaries, therefore providing more active centers. At the same time, nanopores can promote the release of gas products and improve the stability of the catalyst. In addition, the multiple heterogeneous interfaces can modify the electronic structure of the catalyst, and the mixed-valence states of Ni, Mo, and Co further enhance the catalytic performance (Fig. 8j). The optimized MoS2/NiCoS catalyst exhibited overpotentials of 189 mV for HER and 290 mV for OER at 10 mA cm−2 in 1 M KOH solution. In particular, the MoS2/NiCoS catalyst only requires a voltage of 1.50 V to derive a current density of 10 mA cm−2 in a two-electrode electrolyzer (Fig. 8k).

Si et al. used Co3[Fe(CN)6]2 nano cubes as self-sacrificing templates for the in situ growth of a MoS2 nanosheet, during which process Co and Fe are successfully doped into MoS2 (Fe, Co–MoS2). The doping of Co and Fe not only create additional catalytic sites, but also could regulate the electronic structure of MoS2.89 The prepared catalyst retains the cubic geometric structure of the self-sacrificial template, and provides the comprehensive advantages of a high surface area, short charge transfer path, and easily accessible active sites. When used as bifunctional catalysts for water electrolysis, Co,Fe–MoS2 shows excellent catalytic activity with a low cell potential of 1.49 V at a current density of 10 mA cm−2. Adarakatti et al. synthesized a MoS2/CuNi2S4 composite catalyst supported on graphene for HER.90 By introducing transition metal (Cu and Ni) dopants, the adsorption affinity of the active edge to the intermediate hydrogen protons can be increased. Through detailed XPS analysis, it was proved that by anchoring the active center on the graphene framework, the exposure of the unsaturated S edge active sites can be increased, thereby significantly improving the intrinsic hydrogen evolution reaction activity of the MoS2/CuNi2S4 nanosheets.

As illustrated above, nowadays, most of the MSx–MoS2 heterostructure composites focus on Co and Ni-based sulfides, probably due to the high OER activity of cobalt sulfides and nickel sulfides. However, it should be noted that from the viewpoint of commercial application, Fe and Mn-based sulfides are more promising due to their low toxicity and low cost. Therefore, more efforts should be devoted toward developing these potential catalysts.

5. Self-supporting catalysts

As stated above, non-self-supporting catalysts perform well at small current densities and short-term cycles. However, in industrial applications, high current densities and long working-life (such as thousands of hours) are required, and the performance of non-self-supporting catalysts is not satisfactory.91 On the one hand, non-self-supporting catalysts require a polymeric binder to bond it on the working electrode.92 This process inevitably reduces the electrochemically accessible specific surface area, and inhibits the mass and charge transfer rate in the electrochemical process.93 On the other hand, due to the limited cohesive force of the binder, the electrocatalysts may peel off from the working electrode under high current densities,94 resulting in unsatisfactory long-term stability.

Compared with conventional coating catalysts, self-supporting catalysts that are in situ grown on conductive substrates can effectively avoid these defects faced by non-self-supporting catalysts.95 First, due to the strong adhesion of catalysts with the substrates, the self-supporting catalyst will not peel off even if it is cycled under high voltage for a long time. Second, the self-supporting catalysts are usually of nanowire, nanorod, nanosheet or nano bundle morphology, which can expose more active sites and facilitate gas release from its surface. Third, the conductive substrate can further speed up the charge transfer rate.96–98 Therefore, the MoS2-based self-supporting catalysts have attracted more wide research interests in recent years.99,100.

5.1 Self-supporting CoSx–MoS2 composites

Benefiting from the existence of heterogeneous interfaces between MoS2 with CoSx and the excellent conductivity provided by conductive substrate, the designed MoS2–CoSx catalyst in situ grown on a conductive substrate can exhibit excellent bifunctional catalytic performance for HER and OER.101–104 Lu et al. used the Evans–Showell-type polyoxometalate (POM) Co3[Co2Mo10O38H4] as the precursor to construct a CoS2/MoS2 heterogeneous interface through the precise engineering of CoMoS active sites at the atomic level (Fig. 9a).105 The catalytic activity of CoMoS was investigated by density functional theory calculations (Fig. 9b and c). Compared with pure MoS2, which has inferior activity for alkaline HER reaction with a high water dissociation energy barrier, the introduction of Co enables the water decomposition to proceed with a much lower potential energy barrier on the Co site and Co–S site of CoMoS. IN particular, the S site located on the Co edge shows an extremely optimized adsorption energy for both OH and H groups (−0.17 eV). That is, the CoS2/MoS2 heterogeneous interface can promote the HER activity in alkaline media by reducing the water dissociation energy barrier on the Co edge or S atom located at the Co edge.
image file: d0ta10815d-f9.tif
Fig. 9 (a) Schematic representation for the preparation of Co5Mo10Sx/CC. (b and c) Gibbs free energy diagrams for alkaline HER on the possible active sites of MoS2 and CoMoS catalysts. The optimized adsorption surface structures at different stages are shown in the inset. The S, Mo, Co, H, and O atoms are represented by orange, green, blue, white, and red spheres, respectively. Reproduced with permission.105 Copyright 2020, Elsevier. (d) Brief illustration of the synthetic process and structure of the MoS2/CoS2 hybrid and brief illustration of the heteroepitaxy-like growth of MoS2 on CoS2, where yellow balls, blue balls, and red balls respectively represent the S, Mo, and Co elements. (e) XPS survey spectra of MoS2/CoS2, CoS2, and MoS2. (f) Raman spectra of MoS2/CoS2, MoS2, and CoS2. (g) Schematic diagram for the band structure of CoS2 and n-type MoS2 before contact, and the ohmic contact and the charge transfer between metallic CoS2 and n-type MoS2. Reproduced with permission.106 Copyright 2019, American Chemistry Society.

Huang et al. also studied the heterogeneous interface effects between MoS2 and CoS2. The author synthesized a 3D MoS2/CoS2 composite with a shell/core heterostructure that was in situ grown on carbon fiber paper (Fig. 9d).106 According to XPS and Raman analysis (Fig. 9e and f), compared with pure MoS2 and CoS2, the binding energy and Raman vibration peak of the MoS2 phase in MoS2/CoS2 composites shifted negatively, while the binding energy and the Raman vibration peak of the CoS2 phase shifted positively, demonstrating the electron transfer from CoS2 to MoS2. After the electron transfer in the MoS2/CoS2 heterogeneous interface, the CoS2 side forms an electrophilic region, while the MoS2 side forms a nucleophilic region. Due to the difference in electronegativity, for polar water molecules, CoS2 and MoS2 will attract negative O and positive H, respectively. This interaction will promote the cleavage of the O–H bond, thereby promoting water splitting. In addition, according to semiconductor physics, if the work function of the metal is smaller than that of the n-type semiconductor, the metal–semiconductor contact will be an ohmic contact with no obvious junction resistance, so that electrons can be smoothly transferred through the MoS2/CoS2 interface (Fig. 9g).

The optimization of the nanostructure can increase the specific surface area and expose more active sites for electrocatalysts.107–109 Shi et al. prepared a flower-like CoS2/MoS2 hetero-nanosheet array via in situ sulfurization of the CoMoO4 nanosheet array on NF (Fig. 10a).110 The usage of the CoMoO4 nanosheet array as a precursor enabled the well-maintained 3D network structure with high surface area and excellent conductivity after sulfurization (Fig. 10b). The flower-like structure can not only expose more heterogeneous interfaces, but also facilitate the contact of active sites with the electrolyte. Benefiting from the structural advantages and heterogeneous interface effects, the CoS2/MoS2 hetero-nanosheet array showed high catalytic performance for HER, with a low overpotential of 50 mV at a current density of 10 mA cm−2 in 1 M KOH (Fig. 10c).


image file: d0ta10815d-f10.tif
Fig. 10 (a) Schematic illustration of the preparation of CoMoO4 and in situ sulfurization to CoS2/MoS2. (b) Nyquist plots of acid-treated Ni foam and CoS2/MoS2 in 1.0 M KOH. (c) Polarization curves of CoS2/MoS2. Reproduced with permission.110 Copyright 2020, Royal Society of Chemistry. (d) Schematic superhydrophilic/superaerophobic design of the CoMoSx/NF electrocatalysts for overall water splitting. (e) Overall water splitting polarization curves of CoMoSx/NF, MoSx/NF, IrO2–Pt/C and pristine NF electrodes at a scan rate of 5 mV s−1, respectively. (f) Catalytic stability of CoMoSx/NF for overall water splitting at 500 mA cm−2 for 100 hours. Reproduced with permission.113 Copyright 2020, Wiley-VCH.

The water splitting reaction is a process of producing gas; if the generated bubbles cannot be released from the electrode in time, the active sites would be blocked and also impede the diffusion of the electrolyte, resulting in the poor contact between the catalyst and electrolyte.111,112 Compared with non-self-supporting catalysts, self-supporting catalysts more easily achieve surface hydrophilic/hydrophobic engineering.96 Shan et al. reported the preparation of superhydrophilic/superaerophobic amorphous CoMoSx chalcogel catalysts supported on nickel foam (NF) for overall water splitting (Fig. 10d).113 The prepared CoMoSx/NF shows superhydrophilicity, which facilitates the penetration of electrolytes, and superhydrophobicity, which promotes the fast release of the evolving gas bubbles. Moreover, the introduction of Co creates abundant active sites and defects. By employing CoMoSx/NF as bifunctional electrocatalysts, the overall water splitting device delivered 500 mA cm−2 current density at a low voltage of 1.89 V and without decay in performance for 100 hours (Fig. 10e and f).

Although self-supporting composites exhibit higher catalytic activity, it should be noted that the mass loading of the self-supporting catalysts is much higher that of the non-self-supporting catalysts, and few reports in the literature compared the mass activity of the catalysts. For future investigations, it is better to report the mass activity of the catalysts, rather than simply the area activity for self-supporting catalysts.

5.2 Self-supporting NiSx–MoS2 composites

Nickel foam (NF) possesses 3D porous structures and excellent electrical conductivity, which can not only facilitate the good contact between the electrolyte and catalysts, but also reduce the charge transport resistance of the material in the electrochemical reaction.114,115 In addition, nickel foam can serve as a Ni source in the material preparation process to prepare self-supporting electrocatalysts.116–118 Wu et al. synthesized three-dimensional curved nanosheets of Mo-doped Ni3S2 grown on NF through a one-step hydrothermal method.119 The Mo doping can change the morphology of the catalyst from an upright rectangular to wavy nanosheet interweave. The curved and patulous morphology could enable more abundant catalytic active sites for OER and HER (Fig. 11a). Yang et al. used ammonium molybdate and nickel foam as raw materials to synthesize MoS2–Ni3S2 heterogeneous rods supported on nickel foam through the one-pot hydrothermal method (Fig. 11b).120 MoS2–Ni3S2 electrocatalysts displayed superior performance for HER and OER, which achieved low overpotentials of 98 and 290 mV at 10 mA cm−2, respectively. As for the overall water splitting, a cell voltage of only 1.50 V is required for a j of 10 mA cm−2. Such excellent activity can be ascribed to the good exposure of the heterogeneous interface in the hierarchical nanostructure, which is conducive to the chemical adsorption of H- and O-containing intermediates, and the facilitated electron transport along 1D Ni3S2 that is anchored on the NF substrate. Zhang et al. investigated the interface engineering of MoS2/Ni3S2 heterostructures supported on nickel foam.121 The inner Ni3S2 nanoparticles were modified by MoS2 nanosheets on their surfaces to result in abundant MoS2/Ni3S2 hetero interfaces. In alkaline medium, MoS2/Ni3S2 exhibits excellent HER and OER catalytic activity. When MoS2/Ni3S2 is used as a bifunctional catalyst, an alkali electrolyzer delivers a current density of 10 mA cm−2 at a very low cell voltage of 1.56 V. The author calculated the chemisorption free energies of hydrogen (ΔGH) and hydroxide (ΔGOH) on the (101) surface of Ni3S2 and the (002) surface of MoS2 by density functional theory. It was found that compared with the H adsorption energy of the (101) plane of Ni3S2 and the (002) plane of MoS2, H was more easily adsorbed on the Mo–S edge of the Ni-doped MoS2 model, which has the lowest H chemisorption free energy (3.17 eV). Similarly, due to the lowest OH chemisorption free energy (−2.92 eV), oxygen-containing intermediates tend to adsorb on the undercoordinated Ni sites of the Mo-doped Ni3S2 model. In summary, the MoS2/Ni3S2 heterogeneous interface accelerates the decomposition of H2O in alkaline media by reducing the adsorption free energy for intermediate product, and finally, the HER and OER activities of the bimetallic sulfide are enhanced.
image file: d0ta10815d-f11.tif
Fig. 11 (a) Illustration of the designed Mo-doped Ni3S2 nanosheets.119 Copyright 2018, Elsevier. (b) Illustration of the fabrication of the heterostructured MoS2–Ni3S2 HNRs/NF composites. Reproduced with permission.120 Copyright 2017, American Chemistry Society.

The bimetallic sulfide MoS2/NiSx heterogeneous interface can adjust the intermediate species' adsorption energy. However, due to the greater electronegativity of S, the strong interaction between S and the adsorbed H hinders the Tafel/Heyrovsky step in HER. Introducing non-metallic anions to adjust the electron density and catalytic active centers of the bimetallic sulfides is a suitable strategy to improve the Tafel/Heyrovsky step of HER.122–124 Yang et al. synthesized nitrogen anion-decorated NiS/MoS2 heterostructures for HER. The interface between MoS2 and NiS can be clearly seen from the TEM image, indicating the formation of unique NiS/MoS2 hetero-interfaces. Through first-principles density functional theory calculations,125 the author first calculated the effect of N doping on the electron density of N–NiS. As shown in Fig. 12a, the density of states of N–NiS and NiS are continuous at the Fermi level, indicating the intrinsic metallic properties of N–NiS after the introduction of N atoms. Compared with NiS (−4.64 eV), the H2O dissociation free energy of N–NiS (−6.97 eV) is lower. Similarly, compared to MoS2 (0.48 eV), the H2O dissociation free energy of N–MoS2 (−0.06 eV) is also much smaller (Fig. 12b), proving that the introduction of non-metal atoms can enhance the electrocatalytic performance. By calculating the charge density difference, it is shown that the charge polarization of N–NiS and N–MoS2 causes a strong charge redistribution at the heterogeneous interface (Fig. 12c), and the ΔGH of the N–NiS/MoS2 heterojunction is close to 0, which makes it exhibit better HER catalytic performance. In addition, N–NiS/MoS2 exhibits very good stability without a noticeable decrease in the current density after 20 h chronoamperometry measurement, and the morphology of the catalysts is well maintained (Fig. 12d).


image file: d0ta10815d-f12.tif
Fig. 12 (a) Calculated density of states for NiS and N–NiS. The Fermi level is set at 0 eV. (b) Adsorbed H2O dissociation configuration on the surfaces of the MoS2 (002) facet, N–MoS2 (002) facet, NiS (100) facet and N–NiS (100) facet, as well as the corresponding dissociation energies in alkaline solution. Yellow, dark blue, green, light blue, red and white balls represent the S, Ni, Mo, N, O and H atoms, respectively. (c) The charge density difference in the heterostructures of N–NiS/MoS2. The purple and blue isosurfaces represent the charge accumulation and depletion in the space, respectively. (d) Current density versus time (it) curves of N–NiS/MoS2 recorded for 20 h at 30 mA cm−2. Inset: polarization curves before and after the it test.125 Copyright 2019, Royal Society of Chemistry. (e–i) Contact angle of the as-prepared catalysts. The calculated hydrogen adsorption free energy changes for NiS, N–NiS, MoS2, and N–MoS2. Reproduced with permission. (e–i) Contact angle of the as-prepared catalysts: (e) NF, (f) MoS2/NF, (g) Ni3S2/NF, (h) MoS2/Ni3S2/NF, and (i) p-MoS2/Ni3S2/NF. (j and k) Picture of (j) MoS2/Ni3S2/NF and (k) p-MoS2/Ni3S2/NF during the reaction under the same electrocatalytic condition. Reproduced with permission.128 Copyright 2018, American Chemistry Society.

The H2O adsorption on the surface of the catalysts is the first step in the water splitting process, and this step is affected by the wettability of the material surface. Hydrophilic catalysts can adsorb water molecules well. The better the hydrophilicity, the better the compatibility with water; an improved wettability can strongly enhance the water splitting performance.126,127 Li et al. synthesized a superhydrophilic porous MoS2/Ni3S2 catalyst grown on nickel foam (p-MoS2/Ni3S2/NF) through electrodeposition and the solvothermal reaction.128 The author proved the hydrophilicity of the materials and the relationship between hydrophilicity and pore structure through the contact angle test. As shown in Fig. 12e–i, Ni3S2/NF, MoS2/NF and NF all exhibit hydrophobicity, whose contact angle is greater than 90°. Conversely, the contact angle of nonporous MoS2/Ni3S2/NF is 70°, and a water droplet can be quickly absorbed by MoS2/Ni3S2/NF within 0.5 s. As for p-MoS2/Ni3S2/NF, water droplets are completely adsorbed as soon as they touch the surface of the catalyst, demonstrating its superhydrophilicity. During the reaction, a large number of O2 and H2 bubbles gathered in the non-porous MoS2/Ni3S2/NF. In contrast, no large bubbles were observed on the p-MoS2/Ni3S2/NF catalyst. This was probably due to the quick adsorption of water that promoted the squeeze of gas bubbles (Fig. 12j and k).

5.3 Self-supporting MoS2-based trimetallic sulfides composites

MoS2 possesses rich unsaturated S boundaries. Experiments and theoretical calculations have proved that MoS2 is an excellent HER catalyst.129 However, its OER performance under alkaline conditions is not satisfactory, which limits the use of MoS2 as a bifunctional catalyst for water splitting.130 In order to solve this problem, a good strategy is to integrate MoS2 with high active OER catalysts to adjust the electronic structure of the heterogeneous interfaces and accelerate the water dissociation.131 In recent years, it has been proved that NiFe-based composites are excellent OER catalysts.132,133 Therefore, integrating NiFe-based composites with MoS2 can achieved high performance bifunctional catalysts for overall water splitting.

Using FeNi foam as a substrate for the growth of MoS2, Wu et al. synthesized MoS2/Fe5Ni4S8 heterostructures and investigated their performance for water splitting. The Fe5Ni4S8 (FNS) is in situ grown at the interface during the synthesis of MoS2.134 XPS measurements showed that the Mo 3d5/2 and 3d3/2 feature peaks shifted negatively, indicating the strong electronic interactions between FNS and MoS2. The synthesized MoS2/FNS/FeNi foam achieved a current density of 10 mA cm−2 at very low overpotentials of 120 and 204 mV for HER and OER, respectively (Fig. 13a and b). In addition to the high catalytic activity, the MoS2/FNS/FeNi also possesses higher HER stability than that of Pt/C (72.6% vs. 63.2% current retention) and comparable OER stability with IrO2 (Fig. 13c and d). The author clarified the influence of the MoS2/FNS heterogeneous interface on the enhancement for HER and OER through density functional theory calculations. First of all, for HER, the author found that compared to the H adsorption energy (−0.649 eV) on the (103) plane of MoS2, the H adsorption energy on the (103) plane at the S edge sites of FeNi–MoS2 is −1.575 eV, which leads to the tendency of the H protons to be adsorbed on the S boundary sites of the MoS2/FNS heterogeneous interface, thereby enhancing the HER performance. Secondly, for OER, the OH chemisorption energy of the uncoordinated Fe sites on Mo–Fe9S8 is −2.921 eV, which is lower than that of Fe9S8 (−0.940 eV). Similarly, for Mo–Ni9S8, the OH chemisorption energy at the uncoordinated Ni site is −2.438 eV, which is lower than that of Ni9S8 (−1.839 eV). These results prove that the MoS2/FNS heterogeneous interface facilitates the absorption of H atoms and OH groups, consequently enhancing the performances for HER and OER. As shown in Fig. 13e, the author gives the mechanisms of HER and OER on MoS2/FNS.


image file: d0ta10815d-f13.tif
Fig. 13 (a and b) Polarization curves of the MoS2/FNS/FeNi foam, FNS/FeNi foam, FeNi foam, and MoS2/FTO in 1 M KOH for HER and OER, respectively. (c) Chemisorption models and corresponding adsorption energy of H and OH on the surfaces of MoS2, Fe9S8, Ni9S8, MoS2/FNS (Fe, Ni–MoS2 model), MoS2/Fe9S8 (Mo–Fe9S8 model), and MoS2/Ni9S8 (Mo–Ni9S8 model) heterostructures. (d) Current density versus time (it) curves of the MoS2/FNS/FeNi foam and MoS2/FNS/FeNi foil for HER and OER. (e) Illustration for the fabrication of (Ni, Fe)S2@MoS2 heterostructures. Reproduced with permission.134 Copyright 2018, Wiley-VCH. (f) Illustration for the fabrication of (Ni, Fe)S2@MoS2 heterostructures. (g) In situ Raman spectra of (Ni, Fe)S2@MoS2 at different potentials at range of 2000–2700 cm−1, (h) In situ Raman spectra at 2000–2700 cm−1 for (Ni, Fe)S2@MoS2, (Ni, Fe)S2 and MoS2. (i) LSV curves of two-electrode cell assembled by different materials (inset: the optical image of two-electrode setup). (j) Stability test of different electrolytes for (Ni, Fe)S2@MoS2. Reproduced with permission.135 Copyright 2019, Elsevier.

Liu et al. synthesized (Ni, Fe)S2@MoS2 heterostructures for HER and OER through a two-step method. First, NiFe–LDH nanosheets supported on carbon fiber paper were synthesized by hydrothermal method. Then, thioacetamide and sodium molybdate were added to generate (Ni, Fe)S2 nano-boxes covered with layers of MoS2 nanosheets (Fig. 13f).135 The author recorded the formation process of S–Hads through in situ Raman spectra and revealed that the heterogeneous interface formed between trimetallic sulfides is favored by the adsorption of H and the formation of S–Hads, therefore boosting its overall water splitting performance (Fig. 13g and h). Using (Ni, Fe)S2@MoS2 to assemble a water splitting device, the electrolyzer only needs 1.56 V to achieve a current density of 10 mA cm−2 (Fig. 13i). In addition, by adjusting the interface between (Ni, Fe)S2@MoS2 and the electrolyte through introducing S ions into the electrolyte, the leaching of atomic S in the sulfides can be effectively relieved, and the stability of the catalysts can be significantly improved (Fig. 13j).

NiCo-based composites also showed excellent OER activity.136,137 Yang et al. synthesized a hierarchical co-assembly of interacting MoS2 and Co9S8 nanosheets attached on Ni3S2 nanorod arrays, which are supported on nickel foam by hydrothermal method using nickel foam as the substrate and Ni source, thiourea as the S source, and cobalt sulfate and sodium molybdate as the Co and Mo sources, respectively (Fig. 14a).138 When the ratio of Mo[thin space (1/6-em)]:[thin space (1/6-em)]Co is 1[thin space (1/6-em)]:[thin space (1/6-em)]3, the obtained CoMoNiS–NF-31 shows that Co9S8 and MoS2 nanosheets are directly grown on Ni3S2 nanorods, and there is no excess nanosheet to form an aggregation. This structure can make full use of the active centers. In addition, Co9S8, MoS2 and Ni3S2 completely cover the original NF surface, which improves the durability of the 3D frame structure, making CoMoNiS–NF-31 display an outstanding balance between the catalytic activity and stability, according to the activity-stability factor. The author investigated the influence of the MoS2/Co9S8 heterogeneous interface on the OER catalytic activity through density functional theory calculations (Fig. 14b–d). The two structures of the MoS2 layer parallel and perpendicular to the Co-terminated Co9S8 interface were chosen to study the charge transfer in OER. The results show that when Co9S8 is combined with MoS2, the Fermi level will decrease, and electrons will be transferred from Co9S8 to MoS2 through sulfur atoms at the Co-terminated interface (Fig. 14e). By calculating the O*, OH* and OOH* intermediates at different surface positions, the author estimates that the Gibbs free energy of the four electrochemical reaction steps occurred in OER on the Co9S8 surface in the Co9S8/MoS2 (Fig. 14f). It was found that when Co9S8 transfers electrons to MoS2, the binding energy of the OER adsorption species changes and the overpotentials decrease, indicating that the OER activity of Co9S8 strongly depends on the existence of heterogeneous interfaces with other transition metal sulfides. The CoMoNiS–NF-31 demonstrates the best electrocatalytic activity, giving ultralow overpotentials (113, 103, and 117 mV for HER and 166, 228, and 405 mV for OER) to achieve a current density of 10 mA cm−2 in alkaline, acidic, and neutral electrolytes, respectively.


image file: d0ta10815d-f14.tif
Fig. 14 (a) Schematic illustration of the synthesis and growth of hierarchical CoMoNiS–NF-xy composites via one-pot hydrothermal method. (b and c) DFT-optimized structures of the MoS2/Co9S8 interfaces: (b) MoS2 (001) parallel to Co9S8 (001) and (c) MoS2 (001) perpendicular to Co9S8 (001). Atom colors: Mo, black; Co, blue; and S, yellow. Co-termination at the interface results in the most stable structure. (d) Top view of the Co9S8 (001) surface sites used for DFT calculations of the adsorbed O*, OH*, and OOH* species. (e) Calculated densities of the electronic states for the MoS2/Co9S8 interfaces with Co termination and with Co9S8. (f) Gibbs free energy changes for the four steps of OER at 0 V vs. standard hydrogen electrode. Reproduced with permission.138 Copyright 2019, Royal Society of Chemistry.

Mo and W have similar electronic structures. According to theoretical studies, the introduction of foreign metal atoms with similar atomic radius and electronic configuration will inevitably lead to the imbalance of the local Coulomb force, and the derivative disturbance from the rearrangement of atoms will generate additional exposed edge sites. Therefore, the introduction of W will cause changes in the electronic structure of Mo-based materials and improve the overall conductivity and catalytic activity. Zheng et al. synthesized few-layered Mo(1−x)WxS2 hollow nanospheres-modified Ni3S2 heterostructures through the one-step hydrothermal reaction.139 The layered hollow structure of Mo(1−x)WxS2 not only exposes a large number of accessible active sites, but also is conducive to the mass and charge transfer during the reaction. The doping of W establishes a ternary sulfide coupling interface, which enables a strong electronic interaction between Mo(1−x)WxS2 and Ni3S2 to optimize the electronic structure at the interface and improve the conductivity of the ternary sulfide. As a result, the Mo(1−x)WxS2@Ni3S2 heterostructure exhibited good bifunctional performance, with a cell voltage of 1.62 V to deliver a current density of 10 mA cm−2 for water splitting.

6. Summary and outlook

The development of high-efficiency and low-cost HER/OER catalysts is critical to the application of water splitting devices. MoS2 has been proved to possess excellent HER intrinsic activity due to its abundant unsaturated S boundary by theory and experiment. In order to optimize the HER performance of MoS2-based catalysts and enhance their OER performance, constructing multiple heterogeneous interfaces is a good and effective strategy. In this review, from two major aspects of non-self-supporting catalysts and self-supporting catalysts, we reviewed the recent progress on the construction of heterogeneous interfaces between MoS2 and other transitional metal sulfides for water splitting. The preparation process and structural characteristics of each material are introduced, the enhancement effect of the heterogeneous interface on the HER/OER performances of the catalysts are addressed, and the possible catalysis mechanism are discussed to give readers a deep comprehension of the advantages of this kind of electrocatalyst.

As a branch in the development of water splitting catalysts, MoS2 based multi-metal sulfides have received a lot of attention in recent years and outstanding performances were achieved. However, we could not ignore some of the shortcomings and challenges in the current research, and some of the more critical points should be addressed in future investigations.

(i) Insufficient understanding of the reaction mechanism. Although great progress has been made in developing MoS2-transitional metal sulfides for water-slitting, the catalytic mechanism is still not clear. Nowadays, the most frequently proposed mechanism is the electron transfer between the different metal sulfides that promotes the H or OH adsorption. However, it should be noted that the desorption step of the products is still very important and the related investigated is rare. In addition, the interaction between the heterogeneous interfaces composed of multiple metals needs to be studied systematically to clearly explain how the so-called “synergistic effect” occurs. Moreover, the metal sulfides are an unstable species, especially under electrochemical oxidization conditions, like the metal phosphides and metal selenides. It is proposed that the metal sulfides will transform into metal (hydro)oxides during the reaction process. However, this is ignored in most of the current literature. Therefore, it is urgent to clarify the real active sites and catalytic mechanism. The use of theoretical calculations and in situ technology, coupled with the proof of experimental results, will help research studies in better understanding the mystery of water splitting.

(ii) Systematic research is still lacking. Although many of the metal sulfides have been composited with MoS2 to investigate the performance of the catalysts, a systematic approach is still lacking nowadays. It is well known that there are many crystalline phases for metal sulfides. For example, CoS, CoS2, Co3S4 and Co9S8 are common cobalt sulfides and each of them have been composited with MoS2, and many catalytic data have been collected. However, it should be noted that the electrocatalysts are in various nanostructures, morphologies, crystallinity, and combination mode with MoS2. Therefore, we could not compare the performance of the catalysts effectively. One of the major difficulties is that the synthesis of various crystalline-phased metal sulfide–MoS2 composites with the same nanostructures, morphologies, crystallinity is very challenging. In addition, optimizing the ratio of the metal sulfides and MoS2 in the heterostructure to achieve the best performance is rare, and lots of work should be done to reveal this. Theoretical calculations may be a good strategy to selectively predict the best catalysts for saving time and labor costs.

(iii) More effective and cost-effective supports should be developed. For self-supporting catalysts, nowadays, the most widely utilized supports are Ni foam and carbon cloth. For the use of Ni foam supports, it is inevitable to introduce NiSx in the catalysts, which not only limit the diversity of the composites, but also bring difficulties in understanding the catalytic mechanism due to the complicated compositions and interfaces of the catalysts. The utilization of carbon as supports could avoid the above issues. However, the high cost of carbon cloth hinders its widespread applications. Therefore, it is still urgent to develop more diverse and cost-effective supports, especially the supports that involve easy-to-implement hydrophilic/hydrophobic engineering to achieve improved catalytic performance based on the advantaged of the substrates.

(iv) The evaluation criteria need to be improved. At present, most of the evaluations for water splitting catalysts are based on the overpotential under 10 mV cmgeo−2. Although this standard can roughly distinguish the performance of the catalyst, the area of the working electrode is not equivalent to the area used by the catalyst. Moreover, this evaluation standard is based on different loadings and various test conditions. It is also difficult to compare the performance of the catalysts based on different literature studies. Therefore, standard testing conditions should be established for better comparison of the performance of catalysts, and at least the authors should clearly describe the test conditions in their manuscripts. Furthermore, nowadays most of the literature studies are focused on evaluating the performance of the catalysts in aqueous condition in the single cell, where the mass transportation is limited and the oxygen and hydrogen is mixed. For practical application, a membrane-based electrolyzer would be more efficient and the related investigation is very rare. Therefore, it is more significant to construct a membrane-based electrolyzer to evaluate the activity and practical application of the catalysts in the future.

(v) The nanostructures of the catalysts need to be optimized. The structure of the catalyst has a great influence on its electrocatalytic performance. On the one hand, a larger specific surface area means more active sites. On the other hand, an excellent structure can optimize the mass transfer capacity during the catalytic process, leading to better kinetic performances. The water splitting is a process of producing gas, which makes the requirement for the mass transfer capacity of the catalyst higher. However, at the current stage, it is still challenging to effectively construct porous structures, especially hierarchical structures that possess both high specific surface area and fast mass transport. Therefore, more effective strategies should be developed to construct superior nanostructure catalysts for water splitting.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This work has been financially supported by the National Natural Science Foundation of China (No. 51972293, 51772039, 21902189 and 21901264), Key Scientific Research Projects of Universities in Henan Province (21A150062) and Young Backbone Teacher of Zhongyuan University of Technology.

Notes and references

  1. M. Simoes, S. Baranton and C. Coutanceau, Appl. Catal. B Environ., 2010, 93, 354–362 CrossRef CAS .
  2. H. Wen, G.-X. Cao, M.-H. Chen, Y.-P. Qiu, L.-Y. Gan and P. Wang, Appl. Catal. B Environ., 2020, 268, 118388 CrossRef CAS .
  3. T. Q. Guo, Y. Z. Song, Z. T. Sun, Y. H. Wu, Y. Xia, Y. Y. Li, J. H. Sun, K. Jiang, S. X. Dou and J. Y. Sun, J. Energy Chem., 2020, 42, 34–42 CrossRef .
  4. C. C. Liu, T. Gong, J. Zhang, X. R. Zheng, J. Mao, H. Liu, Y. Li and Q. Y. Hao, Appl. Catal. B Environ., 2020, 262, 118245 CrossRef CAS .
  5. F. Safari and I. Dincer, Energy Convers. Manag., 2020, 205, 112182 CrossRef CAS .
  6. D. Zheng, J. Li, S. Ci, P. Cai, Y. Ding, M. Zhang and Z. Wen, Appl. Catal. B Environ., 2020, 277, 119178 CrossRef CAS .
  7. Y. Zhao, N. Jia, X.-R. Wu, F.-M. Li, P. Chen, P.-J. Jin, S. Yin and Y. Chen, Appl. Catal. B Environ., 2020, 270, 118880 CrossRef CAS .
  8. C. L. Hu, L. Zhang and J. L. Gong, Energy Environ. Sci., 2019, 12, 2620–2645 RSC .
  9. M. Sangareswari and M. M. Sundaram, Appl. Water Sci., 2017, 7, 1781–1790 CrossRef CAS .
  10. G. Q. Zhao, K. Rui, S. X. Dou and W. P. Sun, Adv. Funct. Mater., 2018, 28, 1803291 CrossRef .
  11. J. Zhu, L. S. Hu, P. X. Zhao, L. Y. S. Lee and K. Y. Wong, Chem. Rev., 2020, 120, 851–918 CrossRef CAS .
  12. L. Xu, Y. Q. Zou, Z. H. Xiao and S. Y. Wang, J. Energy Chem., 2019, 35, 24–29 CrossRef .
  13. A. Muthurasu, V. Maruthapandian and H. Y. Kim, Appl. Catal. B Environ., 2019, 248, 202–210 CrossRef CAS .
  14. Z. P. Lin, B. Lin, Z. P. Wang, S. G. Chen, C. W. Wang, M. Y. Dong, Q. Gao, Q. Shao, T. Ding, H. Liu, S. D. Wu and Z. H. Guo, ChemCatChem, 2019, 11, 2217–2222 CrossRef CAS .
  15. J. D. Wu, D. P. Wang, S. Wan, H. L. Liu, C. Wang and X. Wang, Small, 2020, 16, 1900550 CrossRef CAS .
  16. Z. Li, Z. Z. Jiang, W. Y. Zhu, C. C. He, P. Wang, X. Wang, T. X. Li and L. Tian, Appl. Surf. Sci., 2020, 504, 144368 CrossRef CAS .
  17. P. Zhang, X. F. Lu, J. W. Nai, S. Q. Zang and X. W. Lou, Adv. Sci., 2019, 6, 1900576 CrossRef .
  18. Z. P. Liu, L. Zhao, Y. H. Liu, Z. C. Gao, S. S. Yuan, X. T. Li, N. Li and S. D. Miao, Appl. Catal. B Environ., 2019, 246, 296–302 CrossRef CAS .
  19. C. Wang, J. Z. Huang, J. Y. Chen, Z. X. Xi and X. L. Deng, Front. Chem., 2019, 7, 118435 Search PubMed .
  20. F. Yu, L. Yu, I. K. Mishra, Y. Yu, Z. F. Ren and H. Q. Zhou, Mater. Today Phys., 2018, 7, 121–138 CrossRef .
  21. T. Yoon and K. S. Kim, Adv. Funct. Mater., 2016, 26, 7386–7393 CrossRef CAS .
  22. C. Wang, X. Shao, J. Pan, J. Hu and X. Xu, Appl. Catal. B Environ., 2020, 268, 118435 CrossRef CAS .
  23. X. Xin, Y. Song, S. Guo, Y. Zhang, B. Wang, J. Yu and X. Li, Appl. Catal. B Environ., 2020, 269, 118773 CrossRef CAS .
  24. Q. Qin, L. Chen, T. Wei and X. Liu, Small, 2019, 15, 1803639 CrossRef .
  25. H. Xu, Y. Jiao, S. Li, H. Meng, J. Wu, X. Shi, Z. Du, R. Wang and G. Tian, Int. J. Hydrogen Energy, 2020, 45, 13149–13162 CrossRef CAS .
  26. X. Liu, P. Wang, Q. Zhang, B. Huang, Z. Wang, Y. Liu, Z. Zheng, Y. Dai, X. Qin and X. Zhang, Appl. Surf. Sci., 2018, 459, 422–429 CrossRef CAS .
  27. X. Wang, J. Wang, X. Zhang, Q. Tian, M. Liu, N. Cai, Y. Xue, W. Chen, W. Li and F. Yu, ChemCatChem, 2019, 11, 1354–1361 CrossRef CAS .
  28. Z. Liu, L. Zhao, Y. Liu, Z. Gao, S. Yuan, X. Li, N. Li and S. Miao, Appl. Catal. B Environ., 2019, 246, 296–302 CrossRef CAS .
  29. S. Anantharaj, S. R. Ede, K. Karthick, S. Sam Sankar, K. Sangeetha, P. E. Karthik and S. Kundu, Energy Environ. Sci., 2018, 11, 744–771 RSC .
  30. P. Wang, Y. Sheng, F. Wang and H. Yu, Appl. Catal. B Environ., 2018, 220, 561–569 CrossRef CAS .
  31. Y. Jiao, Y. Zheng, M. T. Jaroniec and S. Z. Qiao, Chem. Soc. Rev., 2015, 44, 2060–2086 RSC .
  32. M. Hamdani, R. N. Singh and P. Chartier, Int. J. Electrochem. Sci., 2010, 5, 556–577 CAS .
  33. G. Wu, A. Santandreu, W. Kellogg, S. Gupta, O. Ogoke, H. G. Zhang, H. L. Wang and L. M. Dai, Nano Energy, 2016, 29, 83–110 CrossRef CAS .
  34. X. M. Li, X. G. Hao, A. Abudula and G. Q. Guan, J. Mater. Chem. A, 2016, 4, 11973–12000 RSC .
  35. J. H. Wang, W. Cui, Q. Liu, Z. C. Xing, A. M. Asiri and X. P. Sun, Adv. Mater., 2016, 28, 215–230 CrossRef CAS .
  36. S. Anantharaj, S. R. Ede, K. Sakthikumar, K. Karthick, S. Mishra and S. Kundu, ACS Catal., 2016, 6, 8069–8097 CrossRef CAS .
  37. Q. Ding, B. Song, P. Xu and S. Jin, Chem, 2016, 1, 699–726 CAS .
  38. M. S. Balogun, Y. C. Huang, W. T. Qiu, H. Yang, H. B. Ji and Y. X. Tong, Mater. Today, 2017, 20, 425–451 CrossRef CAS .
  39. N. K. Chaudhari, H. Jin, B. Kim and K. Lee, Nanoscale, 2017, 9, 12231–12247 RSC .
  40. J. Bai, T. Meng, D. Guo, S. Wang, B. Mao and M. Cao, ACS Appl. Mater. Interfaces, 2018, 10, 1678–1689 CrossRef CAS .
  41. Y. Wang, W. Kang, D. Cao, M. Zhang, Z. Kang, Z. Xiao, R. Wang and D. Sun, J. Mater. Chem. A, 2018, 6, 4776–4782 RSC .
  42. X. Wang, J. Wang, X. Zhang, Q. Tian, M. Liu, N. Cai, Y. Xue, W. Chen, W. Li and F. Yu, ChemCatChem, 2019, 11, 1354–1361 CrossRef CAS .
  43. Y. Liu, S. Jiang, S. Li, L. Zhou, Z. Li, J. Li and M. Shao, Appl. Catal., B, 2019, 247, 107–114 CrossRef CAS .
  44. F. Du, L. Shi, Y. Zhang, T. Li, J. Wang, G. Wen, A. Alsaedi, T. Hayat, Y. Zhou and Z. Zou, Appl. Catal., B, 2019, 253, 246–252 CrossRef CAS .
  45. S. Deng, Y. Zhong, Y. Zeng, Y. Wang, X. Wang, X. Lu, X. Xia and J. Tu, Adv. Sci., 2018, 5, 1700772 CrossRef .
  46. M. Ramos, G. Berhault, D. A. Ferrer, B. Torres and R. R. Chianelli, Catal. Sci. Technol., 2012, 2, 164–178 RSC .
  47. H. Zhu, J. Zhang, R. Yanzhang, M. Du, Q. Wang, G. Gao, J. Wu, G. Wu, M. Zhang, B. Liu, J. Yao and X. Zhang, Adv. Mater., 2015, 27, 4752–4759 CrossRef CAS .
  48. S. H. Yang, S.-K. Park, J. K. Kim and Y. C. Kang, J. Mater. Chem. A, 2019, 7, 13751–13761 RSC .
  49. Y. Li, X. Wu, H. Zhang and J. Zhang, Nanoscale, 2019, 11, 15763–15769 RSC .
  50. V. Ganesan, S. Lim and J. Kim, Chem.–Asian J., 2018, 13, 413–420 CrossRef CAS .
  51. H. Li, X. Qian, C. Xu, S. Huang, C. Zhu, X. Jiang, L. Shao and L. Hou, ACS Appl. Mater. Interfaces, 2017, 9, 28394–28405 CrossRef CAS .
  52. Y. Guo, J. Tang, H. Qian, Z. Wang and Y. Yamauchi, Chem. Mater., 2017, 29, 5566–5573 CrossRef CAS .
  53. X. Lei, K. Yu, H. Li and Z. Zhu, Electrochim. Acta, 2018, 269, 262–273 CrossRef CAS .
  54. J. Zhang, J. Fang, J. Han, T. Yan, L. Shi and D. Zhang, J. Mater. Chem. A, 2018, 6, 15245–15252 RSC .
  55. T.-T. Chen, R. Wang, L.-K. Li, Z.-J. Li and S.-Q. Zang, J. Energy Chem., 2020, 44, 90–96 CrossRef .
  56. Y. Zhu, L. Song, N. Song, M. Li, C. Wang and X. Lu, ACS Sustainable Chem. Eng., 2019, 7, 2899–2905 CrossRef CAS .
  57. J. Sun, Z. Huang, T. Huang, X. Wang, X. Wang, P. Yu, C. Zong, F. Dai and D. Sun, ACS Appl. Energy Mater., 2019, 2, 7504–7511 CrossRef CAS .
  58. Y. Guo, T. Park, J. W. Yi, J. Henzie, J. Kim, Z. Wang, B. Jiang, Y. Bando, Y. Sugahara, J. Tang and Y. Yamauchi, Adv. Mater., 2019, 31, 1807134 CrossRef .
  59. Z. Chen, X. Duan, W. Wei, S. Wang and B.-J. Ni, J. Mater. Chem. A, 2019, 7, 14971–15005 RSC .
  60. X. Zheng, X. Han, Y. Zhang, J. Wang, C. Zhong, Y. Deng and W. Hu, Nanoscale, 2019, 11, 5646–5654 RSC .
  61. J. Li and G. Zheng, Adv. Sci., 2017, 4, 1600380 CrossRef .
  62. L. Zeng, K. Sun, X. Wang, Y. Liu, Y. Pan, Z. Liu, D. Cao, Y. Song, S. Liu and C. Liu, Nano Energy, 2018, 51, 26–36 CrossRef CAS .
  63. L. An, J. Feng, Y. Zhang, R. Wang, H. Liu, G.-C. Wang, F. Cheng and P. Xi, Adv. Funct. Mater., 2019, 29, 1805298 CrossRef .
  64. X. Xu, W. Zhong, L. Zhang, G. Liu and Y. Du, J. Colloid Interface Sci., 2019, 556, 24–32 CrossRef CAS .
  65. X.-Y. Zhang, B.-Y. Guo, Q.-W. Chen, B. Dong, J.-Q. Zhang, J.-F. Qin, J.-Y. Xie, M. Yang, L. Wang, Y.-M. Chai and C.-G. Liu, Int. J. Hydrogen Energy, 2019, 44, 14908–14917 CrossRef CAS .
  66. J. He, Y. Chen and A. Manthiram, Adv. Energy Mater., 2019, 9, 1900584 CrossRef .
  67. X. Huang, X. Xu, C. Li, D. Wu, D. Cheng and D. Cao, Adv. Energy Mater., 2019, 9, 1803970 CrossRef .
  68. L. Wang, T. Guo, S. Sun, Y. Wang, X. Chen, K. Zhang, D. Zhang, Z. Xue and X. Zhou, Catal. Lett., 2019, 149, 1197–1210 CrossRef CAS .
  69. X. Xu, W. Zhong, L. Wu, Y. Sun, T. Wang, Y. Wang and Y. Du, Appl. Surf. Sci., 2018, 428, 1046–1055 CrossRef CAS .
  70. S. Mao, G. Lu and J. Chen, Nanoscale, 2015, 7, 6924–6943 RSC .
  71. G. Li, J. Wang, J. Yu, H. Liu, Q. Cao, J. Du, L. Zhao, J. Jia, H. Liu and W. Zhou, Appl. Catal. B Environ., 2020, 261, 118147 CrossRef CAS .
  72. N. Ullah, W. Zhao, X. Lu, C. J. Oluigbo, S. A. Shah, M. Zhang, J. Xie and Y. Xu, Electrochim. Acta, 2019, 298, 163–171 CrossRef CAS .
  73. P. Kuang, M. He, H. Zou, J. Yu and K. Fan, Appl. Catal. B Environ., 2019, 254, 15–25 CrossRef CAS .
  74. Y.-H. Li, M.-Y. Yi, J.-Y. Li, Z.-R. Tang and Y.-J. Xu, Appl. Catal. B Environ., 2019, 257, 117934 CrossRef CAS .
  75. L. Zhang, Y. Guo, A. Iqbal, B. Li, D. Gong, W. Liu, K. Iqbal, W. Liu and W. Qin, Int. J. Hydrogen Energy, 2018, 43, 1251–1260 CrossRef CAS .
  76. L. Liu, X. Liu and S. Jiao, J. Colloid Interface Sci., 2020, 564, 77–87 CrossRef CAS .
  77. S. Chandrasekaran, L. Yao, L. B. Deng, C. Bowen, Y. Zhang, S. M. Chen, Z. Q. Lin, F. Peng and P. X. Zhang, Chem. Soc. Rev., 2019, 48, 4178–4280 RSC .
  78. J. Xu, Y. Zhu, B. S. Yu, C. J. Fang and J. J. Zhang, Inorg. Chem. Front., 2019, 6, 3510–3517 RSC .
  79. S. H. Yu, Z. Tang, Y. F. Shao, H. W. Dai, H. Y. Wang, J. X. Yan, H. Pan and D. H. C. Chua, ACS Appl. Energy Mater., 2019, 2, 5799–5808 CrossRef CAS .
  80. S. Liu, S. Li, K. Sekar, R. Li, Y. Zhu, R. Xing, K. Nakata and A. Fujishima, Int. J. Hydrogen Energy, 2019, 44, 25310–25318 CrossRef CAS .
  81. Q. Ji, C. Li, J. Wang, J. Niu, Y. Gong, Z. Zhang, Q. Fang, Y. Zhang, J. Shi, L. Liao, X. Wu, L. Gu, Z. Liu and Y. Zhang, Nano Lett., 2017, 17, 4908–4916 CrossRef CAS .
  82. N. S. Mikhaleva, M. A. Visotin, A. A. Kuzubov and Z. I. Popov, J. Phys. Chem. C, 2017, 121, 24179–24184 CrossRef CAS .
  83. C. Du, D. Liang, M. Shang, J. Zhang, J. Mao, P. Liu and W. Song, ACS Sustainable Chem. Eng., 2018, 6, 15471–15479 CrossRef CAS .
  84. S. A. Shah, X. P. Shen, M. H. Xie, G. X. Zhu, Z. Y. Ji, H. B. Zhou, K. Q. Xu, X. Y. Yue, A. H. Yuan, J. Zhu and Y. Chen, Small, 2019, 15, 1804545 CrossRef .
  85. X. P. Dai, K. L. Du, Z. Z. Li, M. Z. Liu, Y. D. Ma, H. Sun, X. Zhang and Y. Yang, ACS Appl. Mater. Interfaces, 2015, 7, 27242–27253 CrossRef CAS .
  86. J. Hu, C. Zhang, P. Yang, J. Xiao, T. Deng, Z. Liu, B. Huang, M. K. H. Leung and S. Yang, Adv. Funct. Mater., 2019, 30, 1908520 CrossRef .
  87. X.-Z. Song, F.-F. Sun, Y.-L. Meng, Z.-W. Wang, Q.-F. Su and Z. Tan, New J. Chem., 2019, 43, 3601–3608 RSC .
  88. C. Qin, A. Fan, X. Zhang, S. Wang, X. Yuan and X. Dai, J. Mater. Chem. A, 2019, 7, 27594–27602 RSC .
  89. C. Si, Y. Wu, Y. Sun, Q. Liu, L. Tang, X. Zhang and J. Guo, Electrochim. Acta, 2019, 309, 116–124 CrossRef CAS .
  90. P. S. Adarakatti, M. Mahanthappa, J. P. Hughes, S. J. Rowley-Neale, G. C. Smith, A. S and C. E. Banks, Int. J. Hydrogen Energy, 2019, 44, 16069–16078 CrossRef CAS .
  91. C. G. Morales-Guio, L. A. Stern and X. Hu, Chem. Soc. Rev., 2014, 43, 6555–6569 RSC .
  92. J. Zhang, Q. Y. Zhang and X. L. Feng, Adv. Mater., 2019, 31, 1808167 CrossRef .
  93. C. Xia, Q. Jiang, C. Zhao, M. N. Hedhili and H. N. Alshareef, Adv. Mater., 2016, 28, 77–85 CrossRef CAS .
  94. H. Sun, X. Xu, Z. Yan, X. Chen, L. Jiao, F. Cheng and J. Chen, J. Mater. Chem. A, 2018, 6, 22062–22069 RSC .
  95. Z. S. Wang, J. D. Shen, J. Liu, X. J. Xu, Z. B. Liu, R. Z. Hu, L. C. Yang, Y. Z. Feng, Z. C. Shi, L. Z. Ouyang, Y. Yu and M. Zhu, Adv. Mater., 2019, 31, 1902228 CrossRef .
  96. H. M. Sun, Z. H. Yan, F. M. Liu, W. C. Xu, F. Y. Cheng and J. Chen, Adv. Mater., 2020, 32, 1806326 CrossRef CAS .
  97. J. L. Liu, D. D. Zhu, Y. Zheng, A. Vasileff and S. Z. Qiao, ACS Catal., 2018, 8, 6707–6732 CrossRef CAS .
  98. Z. Z. Ma, Q. Zhao, J. P. Li, B. Tang, Z. H. Zhang and X. G. Wang, Electrochim. Acta, 2018, 260, 82–91 CrossRef CAS .
  99. M. W. Hu, J. Huang, Q. Z. Li, R. Tu, S. Zhang, M. J. Yang, H. W. Li, T. Goto and L. M. Zhang, J. Alloys Compd., 2020, 827, 154262 CrossRef CAS .
  100. X. L. Zhang, J. J. Shao, W. Huang and X. C. Dong, Sci. China Mater., 2018, 61, 1143–1153 CrossRef CAS .
  101. J. Hou, B. Zhang, Z. Li, S. Cao, Y. Sun, Y. Wu, Z. Gao and L. Sun, ACS Catal., 2018, 8, 4612–4621 CrossRef CAS .
  102. L. Yang, L. Zhang, G. Xu, X. Ma, W. Wang, H. Song and D. Jia, ACS Sustainable Chem. Eng., 2018, 6, 12961–12968 CrossRef CAS .
  103. W. Zhang, X. Zhao, Y. Zhao, J. Zhang, X. Li, L. Fang and L. Li, ACS Appl. Mater. Interfaces, 2020, 12, 10280–10290 CrossRef .
  104. P. C. Wang, L. Wan, Y. Q. Lin and B. G. Wang, Int. J. Hydrogen Energy, 2019, 44, 16566–16574 CrossRef CAS .
  105. Y. Lu, X. Guo, L. Yang, W. Yang, W. Sun, Y. Tuo, Y. Zhou, S. Wang, Y. Pan, W. Yan, D. Sun and Y. Liu, Chem. Eng. J., 2020, 394, 124849 CrossRef CAS .
  106. N. Huang, Y. Ding, S. Yan, L. Yang, P. Sun, C. Huang and X. Sun, ACS Appl. Energy Mater., 2019, 2, 6751–6760 CrossRef CAS .
  107. C. W. Liang, P. C. Zou, A. Nairan, Y. Q. Zhang, J. X. Liu, K. W. Liu, S. Y. Hu, F. Y. Kang, H. J. Fan and C. Yang, Energy Environ. Sci., 2020, 13, 86–95 RSC .
  108. R. Yang, Y. M. Zhou, Y. Y. Xing, D. Li, D. L. Jiang, M. Chen, W. D. Shi and S. Q. Yuan, Appl. Catal. B Environ., 2019, 253, 131–139 CrossRef CAS .
  109. L. L. Huang, D. W. Chen, G. Luo, Y. R. Lu, C. Chen, Y. C. Zou, C. L. Dong, Y. F. Li and S. Y. Wang, Adv. Mater., 2019, 31, 1901439 CrossRef .
  110. M. Shi, Y. Zhang, Y. Zhu, W. Wang, C. Wang, A. Yu, X. Pu and J. Zhai, RSC Adv., 2020, 10, 8973–8981 RSC .
  111. H. Li, S. Chen, Y. Zhang, Q. Zhang, X. Jia, Q. Zhang, L. Gu, X. Sun, L. Song and X. Wang, Nat. Commun., 2018, 9, 2452 CrossRef .
  112. X. Sheng, Z. Liu, R. Zeng, L. Chen, X. Feng and L. Jiang, J. Am. Chem. Soc., 2017, 139, 12402–12405 CrossRef CAS .
  113. X. Shan, J. Liu, H. Mu, Y. Xiao, B. Mei, W. Liu, G. Lin, Z. Jiang, L. Wen and L. Jiang, Angew. Chem., Int. Ed., 2020, 59, 1659–1665 CrossRef CAS .
  114. S. X. Yan, S. H. Luo, J. Feng, P. W. Li, R. Guo, Q. Wang, Y. H. Zhang, Y. G. Liu and S. Bao, Chem. Eng. J., 2020, 381, 122695 CrossRef CAS .
  115. P. Tan, B. Chen, H. R. Xu, W. Z. Cai, W. He and M. Ni, Appl. Catal. B Environ., 2019, 241, 104–112 CrossRef CAS .
  116. F. Z. Sun, G. Wang, Y. Q. Ding, C. Wang, B. B. Yuan and Y. Q. Lin, Adv. Energy Mater., 2018, 8, 1800584 CrossRef .
  117. X. Gu, S. Zheng, X. Huang, H. Yuan, J. Li, M. Kundu and X. Wang, Chem. Commun., 2020, 56, 2471–2474 RSC .
  118. C. Liang, W. Cao, L. Zhou, P. Yang, X. Zhao, P. Zhao, R. Qiu, L. Yang, Q. s. Huang and D. Astruc, ChemCatChem, 2020, 12, 1647–1652 CrossRef CAS .
  119. C. Wu, B. Liu, J. Wang, Y. Su, H. Yan, C. Ng, C. Li and J. Wei, Appl. Surf. Sci., 2018, 441, 1024–1033 CrossRef CAS .
  120. Y. Yang, K. Zhang, H. Lin, X. Li, H. C. Chan, L. Yang and Q. Gao, ACS Catal., 2017, 7, 2357–2366 CrossRef CAS .
  121. J. Zhang, T. Wang, D. Pohl, B. Rellinghaus, R. Dong, S. Liu, X. Zhuang and X. Feng, Angew. Chem., Int. Ed., 2016, 55, 6702–6707 CrossRef CAS .
  122. P. Chen, T. Zhou, M. Chen, Y. Tong, N. Zhang, X. Peng, W. Chu, X. Wu, C. Wu and Y. Xie, ACS Catal., 2017, 7, 7405–7411 CrossRef CAS .
  123. Y. Wu, X. Liu, D. Han, X. Song, L. Shi, Y. Song, S. Niu, Y. Xie, J. Cai, S. Wu, J. Kang, J. Zhou, Z. Chen, X. Zheng, X. Xiao and G. Wang, Nat. Commun., 2018, 9, 1425 CrossRef .
  124. P. Chen, T. Zhou, M. Zhang, Y. Tong, C. Zhong, N. Zhang, L. Zhang, C. Wu and Y. Xie, Adv. Mater., 2017, 29, 1701584 CrossRef .
  125. M. Yang, Y. Jiang, S. Liu, M. Zhang, Q. Guo, W. Shen, R. He, W. Su and M. Li, Nanoscale, 2019, 11, 14016–14023 RSC .
  126. M. Das, N. Jena, T. Purkait, N. Kamboj, A. De Sarkar and R. S. Dey, J. Mater. Chem. A, 2019, 7, 23989–23999 RSC .
  127. Q. Hu, G. M. Li, X. F. Liu, B. Zhu, X. Y. Chai, Q. L. Zhang, J. H. Liu and C. X. He, Angew. Chem., Int. Ed., 2019, 58, 4318–4322 CrossRef CAS .
  128. F. Li, D. Zhang, R.-C. Xu, W.-F. Fu and X.-J. Lv, ACS Appl. Energy Mater., 2018, 1, 3929–3936 CrossRef CAS .
  129. L. Li, Z. D. Qin, L. Ries, S. Hong, T. Michel, J. Yang, C. Salameh, M. Bechelany, P. Miele, D. Kaplan, M. Chhowalla and D. Voiry, ACS Nano, 2019, 13, 6824–6834 CrossRef CAS .
  130. P. Hu, Z. Jia, H. Che, W. Zhou, N. Liu, F. Li and J. Wang, J. Power Sources, 2019, 416, 95–103 CrossRef CAS .
  131. Q. Z. Xiong, Y. Wang, P. F. Liu, L. R. Zheng, G. Z. Wang, H. G. Yang, P. K. Wong, H. M. Zhang and H. J. Zhao, Adv. Mater., 2018, 30, 1801450 CrossRef .
  132. Z. W. Gao, J. Y. Liu, X. M. Chen, X. L. Zheng, J. Mao, H. Liu, T. Ma, L. Li, W. C. Wang and X. W. Du, Adv. Mater., 2019, 31, 1804769 CrossRef .
  133. C. H. Wang, H. C. Yang, Y. J. Zhang and Q. B. Wang, Angew. Chem., Int. Ed., 2019, 58, 6099–6103 CrossRef CAS .
  134. Y. Wu, F. Li, W. Chen, Q. Xiang, Y. Ma, H. Zhu, P. Tao, C. Song, W. Shang, T. Deng and J. Wu, Adv. Mater., 2018, 30, 1803151 CrossRef .
  135. Y. Liu, S. Jiang, S. Li, L. Zhou, Z. Li, J. Li and M. Shao, Appl. Catal., B, 2019, 247, 107–114 CrossRef CAS .
  136. H. Xu, Z. X. Shi, Y. X. Tong and G. R. Li, Adv. Mater., 2018, 30, 1705442 CrossRef .
  137. S. H. Bae, J. E. Kim, H. Randriamahazaka, S. Y. Moon, J. Y. Park and I. K. Oh, Adv. Energy Mater., 2017, 7, 1601492 CrossRef .
  138. 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, J. Am. Chem. Soc., 2019, 141, 10417–10430 CrossRef CAS .
  139. M. Zheng, J. Du, B. Hou and C. L. Xu, ACS Appl. Mater. Interfaces, 2017, 9, 26066–26076 CrossRef CAS .

Footnote

These authors contributed equally to this work.

This journal is © The Royal Society of Chemistry 2021