Recent advances in defect-engineered molybdenum sulfides for catalytic applications

Yunxing Zhao ab, Xiaolin Zheng *c, Pingqi Gao *a and Hong Li *bde
aSchool of Materials, Sun Yat-sen University, Guangzhou 510275, China. E-mail: gaopq3@mail.sysu.edu.cn
bSchool of Mechanical and Aerospace Engineering, Nanyang Technological University, 639798, Singapore. E-mail: ehongli@ntu.edu.sg
cDepartment of Mechanical Engineering, Stanford University, California 94305, USA. E-mail: xlzheng@stanford.edu
dCINTRA CNRS/NTU/THALES, UMI 3288, Research Techno Plaza, 637553, Singapore
eCentre for Micro-/Nano-electronics (NOVITAS), School of Electrical and Electronic Engineering, Nanyang Technological University, 639798, Singapore

Received 25th March 2023 , Accepted 6th June 2023

First published on 5th July 2023


Abstract

Electrochemical energy conversion and storage driven by renewable energy sources is drawing ever-increasing interest owing to the needs of sustainable development. Progress in the related electrochemical reactions relies on highly active and cost-effective catalysts to accelerate the sluggish kinetics. A substantial number of catalysts have been exploited recently, thanks to the advances in materials science and engineering. In particular, molybdenum sulfide (MoSx) furnishes a classic platform for studying catalytic mechanisms, improving catalytic performance and developing novel catalytic reactions. Herein, the recent theoretical and experimental progress of defective MoSx for catalytic applications is reviewed. This article begins with a brief description of the structure and basic catalytic applications of MoS2. The employment of defective two-dimensional and non-two-dimensional MoSx catalysts in the hydrogen evolution reaction (HER) is then reviewed, with a focus on the combination of theoretical and experimental tools for the rational design of defects and understanding of the reaction mechanisms. Afterward, the applications of defective MoSx as catalysts for the N2 reduction reaction, the CO2 reduction reaction, metal–sulfur batteries, metal–oxygen/air batteries, and the industrial hydrodesulfurization reaction are discussed, with a special emphasis on the synergy of multiple defects in achieving performance breakthroughs. Finally, the perspectives on the challenges and opportunities of defective MoSx for catalysis are presented.



Wider impact

Molybdenum sulfide catalysts have gained widespread acceptance in industrial petroleum hydrodesulfurization, and they are expanding to the electrochemical reactions. Recent advances have witnessed defect engineering as a powerful tool to effectively activate structures. Herein, we systematically summarize the recent advances in defective molybdenum sulfides for catalytic applications. Compared to previous articles centered around a single application with limited defect types, this article reviews the various defect types of molybdenum sulfides for a range of catalytic applications, emphasizing the correlation between the defect structure, theoretical mechanism, and performance. Besides elaborating the defects in crystals, this article also fills a gap in similar reviews by systematically addressing defects in both amorphous and clustered molybdenum sulfides. The review is highly interdisciplinary, relating to materials science, electrochemistry, thermal catalysis, heterogeneous catalysis, semiconductor materials, and nanotechnology; so, we anticipate that this review will be of great interest to a wide variety of researchers with extensive backgrounds. Moreover, the focus of this review is on defects in molybdenum sulfide, where defect screening, structural design, fabrication methods, and characterization techniques are important elements in catalysis and materials science, and thus this review contributes to the advancement of materials methodology and the development of defect-based industrial catalysts.

1. Introduction

Molybdenum disulfide (MoS2) has gained great popularity as a well-established industrial catalyst for the petroleum hydrodesulfurization (HDS) reaction for oil upgrading, and it is hard to prevent its aggressive expansion into cutting-edge applications in electrochemical catalysis. With its high abundance (10[thin space (1/6-em)]000 times that of platinum) and excellent hydrogen evolution reaction (HER) catalytic properties, MoS2 has been emerging as one of the most promising HER catalysts to replace platinum (Pt) in proton exchange membrane (PEM) electrolyzers, signifying its potential value in the field of hydrogen energy. Moreover, MoS2 extends its catalytic applications in other reactions such as the nitrogen reduction reaction (NRR), carbon dioxide reduction reaction (CO2RR), oxygen reduction reaction (ORR), oxygen evolution reaction (OER), nitric oxide (NO) reduction reaction, nitrate ion (NO3) reduction reaction, and organic molecule conversion. MoS2 also showcases advantages as an electrode material with catalytic functions in energy storage devices such as metal–sulfur batteries and metal–oxygen/air batteries. Broader applications based on MoSx catalysts are yet to be discovered. Bearing this in mind, this paper will review the catalytic applications of defective MoS2 materials in two-dimension (2D) comprising 2H, 1T, and 3R phases and non-2D (amorphous and cluster phases) morphologies, mainly for the HER, as well as for the NRR, the CO2RR, metal–sulfur batteries, metal–oxygen/air batteries and HDS reactions. Before specifically reviewing its applications in catalysis, we will give a brief overview of the structural classification and catalytic applications of MoSx.

1.1. Introduction of molybdenum sulfide

2D MoS2 crystals are constructed by stacked layers, in which an infinitely extended planar monolayer composed of Mo and S elements binds to adjacent layers with interlayer van der Waals forces. Three crystalline phases including phases 2H, 1T, and 3R are defined by different periodic arrangements within and/or between the layers and their unit cell structures are shown in Fig. 1a, where 1, 2, and 3 layers are displayed in the unit cell for phases 1T, 2H, and 3R, from left to right, respectively.1 For the 2H phase, the adjacent layer is rotated 180° and stacked in a direction perpendicular to the expanded basal plane; and for the 3R phase, the adjacent layer is slightly displaced from each other; while for the 1T phase, the adjacent layer remains unchanged. As summarized in Table 1, 2H and 3R phases possess the same trigonal prismatic coordination but with different hexagonal and rhombohedral symmetries, as well as distinct point and space groups, while the 1T phase has an octahedral coordination with trigonal symmetry along with point group D3d and space group P[3 with combining overline]m1. Furthermore, in terms of stability, the semiconducting 2H phase is dominant in natural minerals, followed by the semiconducting 3R phase and its mixture, however, the metallic 1T phase could not be found in nature on account of its metastable properties in thermodynamics, transforming into 2H or 3R phases when heated above 100 °C or just upon aging.2
image file: d3mh00462g-f1.tif
Fig. 1 Crystal and electronic structures of MoS2. (a) Crystal structures of MoS2: 1T, 2H and 3R unit cell structures. Reproduced with permission.1 Copyright 2016, John Wiley & Sons, Inc. (b) The building [MoS6] motifs of 2H- (top) and 1T-phase (bottom) MoS2 and the 4d orbital splitting of Mo4+. (c) Calculated band structures of bulk 2H and 1T phases. (b and c) Reproduced with permission.8 Copyright 2018, John Wiley & Sons, Inc. (d) Variations of the band gap of multilayer 2H MoS2 sheets with the number of sheets. Reproduced with permission.13 Copyright 2007, American Chemical Society. (e) Total DOS of the pristine 2H MoS2 (black) and the MoS2 with grain boundary (red dashed) and projected DOS of atoms along the grain boundary (blue filled). The grey dashed line indicates the Fermi energy of pristine MoS2 and the light green shaded area indicates the pristine band gap. Reproduced with permission.18 Copyright 2013, Springer Nature. (f) Scanning tunneling microscopy (STM) measured differential conductance dI/dV (left y-axis) and DFT calculated projected d-orbital density of states (d-DOS) on the Mo atom of the S-vacancy (right y-axis) as a function of sample bias for monolayer 2H-phase MoS2 with 12.5% S-vacancy. Reproduced with permission.19 Copyright 2016, Springer Nature. (g) Schematic of strained 2H-phase MoS2 indented by SiO2 nanocones, where the regions on the tips of nanocones show the highest strain, while the areas between nanocones are less strained. The MoS2 energy band gap inversely tracks the strain profile and is spatially modulated, forming ‘artificial atoms’ at the points of peak strain. Reproduced with permission.20 Copyright 2015, Springer Nature.
Table 1 A summary of basic structural information for 2H, 1T, and 3R phases of MoS2, where the bold letters (H, T, and R) and numbers (2, 1, and 3) relate to the abbreviations of 2H, 1T, and 3R, respectively
Crystal phase Coordination Symmetry Layer numbers in unit cell Layer arrangement Point group Space group Conductivity Stability
2H Trigonal prismatic Hexagonal 2 Adjacent layer is rotated by 180° and stacked D 3h (odd layer numbers) P63/mmc (P[6 with combining macron]m2 for monolayer) Semiconductor Stable and dominant in nature
D 3d (even layer numbers)
1T Octahedral Trigonal 1 Adjacent layer remains unchanged D 3d P[3 with combining macron]m1 Metal Metastable
3R Trigonal prismatic Rhombohedral 3 Adjacent layer is slightly displaced from each other C 3v R3m (P[6 with combining macron]m2 for monolayer) Semiconductor Stable in nature


With respect to its electronic structure, different atomic arrangements of the MoS6 unit lead to distinct splitting of Mo 4d orbitals according to crystal field theory. In the case of 2H or 3R phases, the d orbitals of the D3h-MoS6 unit split into three groups including dz2, dxy and dx2y2, dxz and dyz; and two electrons occupy the dz2 orbital (Fig. 1b). In contrast, the d orbitals of the Oh-MoS6 unit in the 1T phase split into triply degenerate t2g (dxy, dxz, dyz) and doubly degenerate eg (dx2y2, dz2) orbitals, with two electrons occupying two out of the three t2g orbitals without considering the Jahn–Teller effect.2–4 Engineering of unoccupied orbitals by injecting extra electrons from metals (e.g., Li/Re/Mn) can be used in a wide range of fields, such as crystal phase tuning, ferromagnetism, photocatalysis, superconductivity, etc.2–7 The band structure is formed by the hybridization of Mo and S states, where the Mo 4d orbital contributes to the upper valence band edge and lower conduction band edge while the S 3p orbital mainly contributes to the valence band. For the 2H phase, the valence band maximum (VBM) is located at the Γ point in the Brillouin zone, but the conduction band minimum (CBM) is located approximately halfway along the ΓK direction (Fig. 1c), which leads to an indirect band gap of 1.29 eV in bulk 2H-phase MoS2. In contrast, the Mo 4d states form a broad band hosting the Fermi level in 1T-phase MoS2, suggesting its metallic behavior.3,8–12 Furthermore, as the number of layers decreases, the band gap increases (Fig. 1d).13 Meanwhile, the valence and conduction band states near the Γ point are strongly influenced by the interlayer interactions because of the large contribution of p orbitals of the S atoms to the electronic states, while the states near the K point are marginally affected because they consist mainly of localized d orbitals of Mo atoms, leading to an indirect-to-direct band gap transition of MoS2 from the multilayer to monolayer,13–15 and the monolayer MoS2 with a direct band gap of ∼1.8 eV can be utilized as an advanced material for optoelectronic applications such as phototransistor and field-effect transistor.16,17

We here classify 2H-, 1T-, and 3R-phase crystalline MoS2 as 2D molybdenum sulfides because they are usually prepared in several layers to ensure a larger specific surface area when used as catalytic materials. We classify amorphous and clustered MoSx as non-2D molybdenum sulfides because they do not consist of atomically thin layered units. We consider a perfect MoS2 crystal with n (n ≥ 1) monolayers stacked according to the standard crystal pattern. In the perfect crystal, the defective structure is absent, e.g., no difference in the number of stacked layers, only an infinitely small ratio of edge positions, without discontinuous layers, no voids, lack of distortion, and the absence of dislocations. In fact, edge defects, atomic vacancies, heteroatom adsorption sites, and other types of defects are usually present in natural MoS2 crystals. From the perspective of controllable defect modulation, we focus here on artificial preparation of defect types, including porous structure formation, atomic doping, designing strained structures, dislocation and distortion formation, atom vacancy formation, alloying, hybridization, phase modulation, unsaturated coordination, bonding regulation, surface decoration, heterointerface formation, etc.

The electronic structure of MoS2 can be tuned by internal defects, such as grain boundaries (Fig. 1e) and S-vacancies (Fig. 1f), both of which can introduce mid-gap states that altering the optical, transport, or adsorption properties.18,19 The electronic structure can be further tuned by external factors, for instance, via applying elastic strain on monolayers to obtain a narrower energy gap and decreased exciton energy with increasing tensile strain (Fig. 1g), providing opportunities for next-generation optoelectronic or photovoltaic technologies.20 In addition, a sizable spin–orbital splitting occurs in monolayer MoS2, which breaks the degeneracy of the valence band maximum (VBM) and the conduction band minimum (CBM), displaying its potential for application in spin electronics and quantum information fields.12

1.2. Catalytic application of molybdenum sulfide

2D MoS2 has already been established as an industrial desulfurization catalyst for gasoline production and is being actively extended to electrocatalytic applications. With 2D MoS2 as an example, we briefly illustrate the electrocatalytic applications of molybdenum sulfide. Electrocatalysis comprises a transport process, which relies on electrons and holes in an external circuit and ions in the solution to transfer electrons/holes to the adsorbent before obtaining the target chemicals through the oxidation reaction at the anode and the reduction reaction at the cathode; and a catalytic process, which improves the conversion efficiency of electrical energy to chemical energy by overcoming the kinetic barrier. With the development in the field of electrocatalysis in the last decade, MoS2 has shown great potential for catalytic applications in the HER, OER, ORR, CO2RR, NRR, nitric oxide (NO) reduction reaction, nitrate ion reduction reaction, organic small molecule conversion, and so on, by virtue of its excellent performance, decent electrical conductivity and stable structure in harsh environments.19,21–27Table 2 summarizes a few representative applications of MoS2 in the HER, OER, ORR, CO2RR, and NRR including half-reaction, equilibrium potentials E0, media, and catalysts. MoS2 has shown excellent catalytic performance for the HER in acidic media, but encounters difficulties in catalyzing the HER and the OER in basic media due to its inferior activity compared to Fe–, Co–, and Ni-based materials. Fortunately, engineering of intrinsic MoS2, such as carbon doping, substitutional doping of Ru, construction of heterostructures, and design of biaxially strained MoS2 nanoshells, can generate competitive activity for the catalytic HER in alkaline media.28–31 Fabrication of quantum dots or construction of hetero-interfaces renders MoS2 enhanced OER activity.21,32 Moreover, it also exhibits remarkable activity towards the CO2RR compared to Ag particles, delivering a much higher current density of 65 mA cm−2 at −0.764 VRHE and an exceptional faradaic efficiency (FE) up to 98%.23 With regard to the NRR, intrinsic MoS2 gives a NH3 yield of 8.08 × 10−11 mol s−1 cm−2 at −0.5 VRHE in 0.1 M Na2SO4 with a FE of 1.17%; and further optimization of the catalytic system, e.g., upgrading electrolytes or making composites, could yield an order of magnitude increase in activity and selectivity.33,34 In addition, defective MoS2 is also very useful when serving as a catalyst for batteries. Being an electrode material, it can catalyze the conversion of Li2Sn (3 ≤ n ≤ 8), Na2Sn (4 ≤ n ≤ 8), Li2O2 and O2/OH, in Li–S, Na–S, Li–O2 and Zn–air batteries, respectively.32,35–37
Table 2 Representative electrocatalytic applications of MoS2 catalysts
Application Half-reaction Media CatalystRef. Synthesis method Potential (vs. RHE) Tafel slope (mV dec−1) Faradaic efficiency (%)
a Not provided items in articles are labeled as N/A.
HER 2H+ + 2e → H2 (E0 = 0 VNHE) 0.5 M H2SO4 MoS2 with S-vacancies19 Plasma-treated MoS2 η = 170 mV@10 mA cm−2 60 N/Aa
2H2O + 2e → H2 + 2OH(E0 = −0.826 VNHE) 1.0 M KOH Carbon doped MoS228 Incompletely sulfurized Mo2C η = 45 mV@10 mA cm−2 46 97
OER 2OH → ½O2 + H2O + 2e (E0 = 0.404 VNHE) 1.0 M KOH MoS2 quantum dots21 Hydrothermal synthesis η = 370 mV@10 mA cm−2 39 N/A
ORR O2 + 2H2O + 4e → 4OH (E0 = 0.404 VSHE) 0.1 M KOH Graphene/N-doped MoS222 CVD deposition 0.716 V@ half-wave potential N/A N/A
CO2RR CO2 + 2H+ + 2e → CO + H2O (E0 = −0.52 VSHE, pH 7) EMIM-BF4 and water Vertically aligned MoS2 nanoflake23 CVD deposition −0.164 V@ onset potential N/A 98
NRR N2 + 2H2O + 6H+ + 6e → 2NH3·2H2O(E0 = 0.092 VNHE) 0.1 M Na2SO4 MoS2 nanosheet on carbon cloth24 Hydrothermal synthesis −0.5 V@8.08 × 10−11 mol s−1 cm−2 N/A 1.17


Many studies have focused on defective MoS2, especially on the methods of defect generation (i.e., direct synthesis of defective MoS2 and creation of defects based on the as-prepared crystals) and on the understanding of the effects of defects on activity and stability, to achieve controlled design, fabrication, and use of defective MoS2 towards catalysis. As a special class of catalytically active sites, defects lead to very different properties from conventional crystals due to their structural diversity, high distortion energy, and complex atomic arrangement, which are significant for catalysis. It should be noted that we focus on reviewing the catalytic applications of defective MoSx materials in 2D (2H, 1T, and 3R phases) and non-2D (amorphous, cluster) morphologies. Fig. 2 summarises the catalytic applications of various defective MoSx materials in the HER, NRR, CO2RR, metal–sulfur batteries, metal–oxygen/air batteries, and HDS reactions, and provides navigation for subsequent discussion. Among these electrocatalysis fields, MoS2 has made the most successful breakthrough in acidic HER applications from our perspective, paving the way for the development of platinum-free catalysts to replace the precious platinum in PEM electrolyzers. We will therefore firstly detail the progress of MoS2-based electrocatalysts for HER application and then we discuss the catalytic performance of defective MoS2 in other electrocatalytic fields, including the NRR, CO2RR, metal–sulfur batteries, and metal–oxygen/air batteries and finally we will briefly review its application in hydrodesulfurization for petroleum upgrading.


image file: d3mh00462g-f2.tif
Fig. 2 Navigation map of various defects in MoSx materials and their catalytic applications.

1.3. Preparation of defective molybdenum sulfide

The preparation method of defects determines their structure and thus plays a key role in the catalytic performance. Various novel defect preparation methods have emerged over a long period of time, and the design ideas behind them are worth reviewing. Controlled preparation methods confer reproducibility to the material, which is crucial for its practical applications. In this section, we provide a brief review of several methods for the preparation of defective molybdenum sulfide based on some representative cases.

The porous structure breaks the continuity of the material and provides a large number of defect sites such as edge and unsaturated coordination atoms. Abundant porous structures can be prepared using a template method, chemical exfoliation method, etc. For instance, Skrabalak and coworkers prepared porous nanostructured molybdenum sulfide using an ultrasonic spray pyrolysis (USP) method.38 Specifically, they obtained a MoS2/SiO2 composite by nebulizing and pyrolyzing a mixture of ammonium tetrathiomolybdate and colloidal silica and then cleaned the SiO2 template with HF to obtain a highly porous MoS2 network with a surface area of up to 250 m2 g−1. Yin and coworkers employed a liquid-ammonia-assisted lithiation (LAAL) strategy to exfoliate 2D-MoS2.39 This process includes the lithiation reaction that occurs when liquid ammonia contacts with lithium, the desulfurization reaction between lithium and MoS2, and the exfoliation of Li-intercalated MoS2 by ultrasonication in water. The sample obtained using this method is mesoporous 1T-phase MoS2 nanosheet (P-1T-MoS2), which can be converted into P-2H-MoS2 by annealing. In addition to the pore structure, P-1T-MoS2 also has a plenty of edge sites and numerous S-vacancies and the contents of S-vacancies gradually increase with the ratio of lithium to MoS2.

For the fabrication of S-vacancies in crystalline MoS2, breakthroughs have been achieved using physical or chemical methods, with the former typically involving direct disruption of the Mo–S bond by kinetic energy and the latter allowing S removal by altering electron transfer or chemical bonding. Plasma treatment techniques have been extensively investigated for the preparation of S-vacancies on the basal plane of MoS2. For instance, Li and coworkers created S-vacancies by exposing CVD (chemical vapor deposition)-synthesized monolayer 2H-phase MoS2 to a mild argon plasma atmosphere, where the plasma was generated by dispersing small RF power in a vacuum chamber, which results in a short mean path for the radicals in the plasma and thus ensures a cold and mild Ar plasma on the MoS2 surface, thus making the desulfurization process more controllable.19 Different vacancy concentrations can be realized by the duration of the treatment. Wu and colleagues used a chemical route to prepare MoS2 with in-plane S-vacancies.40 Commercial bulk MoS2 and Zn powder were mixed by grinding, then dried and degassed under an inert atmosphere and subsequently heated at a high temperature (700 °C). During the heating process, Zn as a reductant reacted with MoS2 to produce Mod+ (d < 4), thereby excluding S from the basal plane and producing defective MoS2 nanoparticles of about 25 nm. By adjusting the ratio of Zn to MoS2, the density of S-vacancies and the electrical conductivity can be tuned simultaneously. Additionally, more chemical pathways for creation of S-vacancies have been investigated because of their advantage in large-scale production of defective MoS2.

Strain in MoS2 introduces many defects, such as dislocations or mismatch of atoms within the layer, folding or warping of the two-dimensional layer structure, resulting in the modulation of the electronic structure as well as the catalytic properties. The generation of strain depends on force effects such as substrate-induced strain,19 lattice mismatch-induced strain,41 and shaping-induced strain.42 For instance, Li and coworkers applied strain to the MoS2 monolayer based on a nanostructured substrate.19 Specifically, the CVD-grown MoS2 monolayer was wet-transferred onto a SiO2 nanocone substrate pre-patterned using nanosphere lithography. After removing PMMA with chloroform, the SiO2 nanocone with transferred MoS2 was soaked in ethylene glycol in vacuum environment, ensuring that both sides of the MoS2 sheet were immersed. The sample was then dried in ambient air to evaporate ethylene glycol, when the capillary force pulled the MoS2 against the SiO2 nanocone. The MoS2 region on the tip of the nanocone is most strained and the region between nanocones is less strained.

Doping defects include substitution doping, interstitial doping, surface doping, etc., which regulate the electronic structure, electrical conductivity, and defect density of catalytic materials through the synergistic effect between elements. Although there are many methods of doping, such as the hydrothermal method, co-precipitation, CVD method, etc., they can be simply divided into in situ doping and post-synthesis doping according to the sequence of the doping procedure. For example, Zhang and colleagues employed a one-step hydrothermal method at 220 °C to prepare Ru, O-co-doped MoS2 using H32Mo7N6O28, CH4N2S, and RuCl3 as precursors, where the ruthenium content was adjusted by the ratio of the feed.43 This method yielded molybdenum sulfide with a nanoflower structure, in which Ru and O are doped into the lattice. Sun and colleagues used a two-step method to achieve post-synthesis doping of molybdenum sulfide.44 Briefly, Mn-doped MoS2 nanosheets (Mn–MoS2) were first synthesized in a hydrothermal reactor at 180 °C using ammonium molybdate, manganese chloride, and thioacetamide as precursors. Subsequently, N and Mn-co-doped MoS2 nanosheets (N, Mn–MoS2) were obtained by treating Mn–MoS2 in a tube furnace under a NH3 atmosphere at 300 °C.

MoS2 with mixed crystalline phases forms an abundant phase heterojunction that differs from single-phase crystals in the electronic structure and increases the density of defect sites, offering an opportunity to reduce the catalytic kinetic barrier and to increase the catalytic activity. Both top-down exfoliation and bottom-up synthetic routes have been extensively studied. For instance, Ke and colleagues prepared MoS2 in the 2H/1T phase using an exfoliation method.45 Briefly, the MoS2 powder was immersed in a n-butyllithium solution, placed in a glove box under an inert atmosphere and washed with hexane after 3 days, then it was moved to an air environment. After dispersion in water, sonication in ice water and collection of the powder, monolayer nanosheets were obtained with 51.7% yield of the 1T-phase, which were transformed from the 2H-phase by the charge donation from Li atoms. Ren and colleagues instead used a one-pot hydrothermal method to synthesize MoS2 in mixed phases.46 They used ammonium molybdate tetrahydrate and thiourea as the molybdenum and sulfur sources, respectively, and performed the hydrothermal reaction at 200 °C. They obtained 1T/2H-MoS2 and 1T/3R-MoS2 when the molar ratios of molybdenum and sulfur were controlled at 1[thin space (1/6-em)]:[thin space (1/6-em)]5 and 1[thin space (1/6-em)]:[thin space (1/6-em)]25 and the solutions occupied 45% and 75% of the reactor volume, respectively.

Intercalation could be considered as an interlayer defect with the function of stabilizing metastable phases (e.g., 1T, 1T′) as well as modulating the properties of the guest. Similarly, both in situ intercalation during synthesis and post-synthesis intercalation based on MoS2 have been investigated. For example, Deng and colleagues prepared (N, PO43−)-MoS2/VG using an in situ route.47 Briefly, aqueous solution of sodium molybdate and thioacetamide and vertical graphene (VG) films were transferred to a Teflon-lined steel autoclave and held at 200 °C for 12 h. Subsequently, an MoS2/VG core/shell array with N doping (28%) and PO43− intercalation (10%) was obtained by annealing the films at 300 °C. Feng and coworkers used a polymer-direct-intercalation (PDI) strategy to construct a MoS2/N-doped carbon (MoS2/NC) heteroaerogel.48 Specifically, PEI (polyethyleneimine) and MoS2 were ultrasonically mixed to form a suspension, during which the PEI molecule with positively charged NH2+ group was adsorbed on or inserted into the surface or the interlayer of the negatively charged MoS2 nanosheets. The suspension was then freeze-dried to obtain aerogel-like 3D MoS2-PEI due to the linkage of PEI molecules between MoS2 nanosheets. Black MoS2/NC was obtained by further annealing MoS2-PEI at 800 °C under an argon atmosphere, in which the heteroaerogel shape was well preserved.

Single atoms anchored on the surface of molybdenum disulfide realize a synergistic catalysis with the carrier by modulating the local electronic structure. Methods such as interfacial reaction adsorption and one-step synthesis are effective options.49,50 For instance, Li and coworkers prepared an FeSA/MoS2 catalyst via a spontaneous reduction reaction of Fe3+ in defect-containing MoS2.49 Specifically, MoS2 nanosheets were firstly synthesized by heating a mixture of ammonium tetrathiomolybdate and N,N-dimethylformamide in a Teflon-lined autoclave at 220 °C. The obtained MoS2 was sonicated in H2O2 solution (3 wt%), during which abundant defects were introduced by H2O2 etching. The sample was then dispersed in an FeCl3 solution before stirring, during which the defect-rich MoS2 induced the reduction of Fe3+ due to its redox properties and individual Fe anchored to its surface. The sample was further dispersed in (3-mercaptopropyl)trimethoxysilane/dichloromethane, where the thiol group of the molecules refilled some S-vacancies and was bound to Mo. Finally, dissociation of the alkyl group of the molecules was realized by heating the sample in H2/Ar at 300 °C. The single-atom Fe (loading of 0.31 wt%) was confirmed by AC HAADF-STEM as well as by X-ray absorption spectroscopy, i.e., each Fe is coordinated with three adjacent S and located on top of Mo.

For the preparation of MoS2-based alloy materials, an in situ preparation technique and a post-synthesis conversion technique are both applicable. Wang and coworkers used a one-step CVD method to synthesize MoS2(1−x)Se2x/SnS2(1−y)Se2y heterojunction nanosheets.51 The solid sources include MoO3/SnSe in the center of the furnace, and S powder at the upstream. The furnace was ramped up to the growth temperature (650 °C) under an N2 flow (200 sccm) and the vapor-phase reaction was performed for 4 min under another N2 flow (10 sccm) to grow both types of layers simultaneously. The AFM images as well as Raman spectra showed that SnS2(1−y)Se2y was fully or partially grown on the MoS2(1−x)Se2x sheets, while Mo–S and Mo–Se bonds coexisted in MoS2(1−x)Se2x. Zhang and colleagues obtained a MoS2−xSex alloy/graphene composite by direct selenization of MoS2/graphene.52 In brief, they mixed hydrothermally synthesized MoS2/graphene with Se powder, then placed the mixture in a tube furnace and heated it at 700 °C for 2 h in a N2 flow. MoS2−xSex alloy exhibited an expanded interlayer spacing and an S[thin space (1/6-em)]:[thin space (1/6-em)]Se ratio of 1.2[thin space (1/6-em)]:[thin space (1/6-em)]0.8.

To prepare amorphous or clustered MoSx, electrodeposition,53 sputtering deposition,54 and wet chemistry methods have been extensively studied.55 For example, Escalera-López and coworkers conducted electrochemical deposition to prepare an amorphous MoSx film on a Si/Ti/Au substrate.53 Deoxygenated solution containing ammonium tetrathiomolybdate and NaClO4, saturated Ag/AgCl, and a Pt mesh were employed as the electrolyte, reference electrode, and counter electrode, respectively. The Pt electrode was encapsulated in a glass vial with a fritted junction to prevent Pt deposition on the MoSx films. AE-MoSx was obtained by anodic electrodeposition at a constant voltage of +0.1 V (versus Ag/AgCl), while CE-MoSx was obtained by cathodic electrodeposition at −1 V (versus Ag/AgCl). The relationship between film the thickness and charge density was established by measuring the film with a surface profilometer. Kibsgaard and colleagues used a wet chemical method to prepare Mo3S132− clusters.55 Ammonium polysulfide solution (25 wt%) was added to ammonium molybdate solution and the mixture was left in an oil bath (96 °C) for 5 days until dark-red crystals precipitated out. After sequential filtering and washing, the product was then heated in hot toluene (80 °C) for several hours to remove excess sulfur to obtain (NH4)2Mo3S13·nH2O.

Putting together, there are often multiple types of defects prepared using a particular method and there are multiple preparation methods for a particular defect type. The above review does not provide a relatively comprehensive coverage of the methods for preparing defective molybdenum sulfide and we will continue our discussion of methods in later sections.

2. Defective two-dimensional molybdenum disulfide for the HER

Hydrogen is considered as an ideal energy source to address the challenges due to environmental pollution and energy shortage. Fossil fuel-based hydrogen production processes generate harmful carbon emissions (7.33–29.33 kg CO2 emissions for H2 kg−1 production at 75% system efficiency), while water electrolysis utilizing electricity converted from wind-, hydro- and solar- power could achieve carbon-free emissions.56 Proton exchange membrane (PEM) electrolyzers with many advantages, including a high operating current density, high-purity pressurized hydrogen, good partial load range (i.e., compatible with intermittent renewable energy sources), high efficiency and portability/compactness,57 rely heavily on scarce catalysts including platinum (Pt) and ruthenium (Ru)/iridium (Ir) for the HER and the OER, respectively. The cost of highly loaded platinum group metals (>3 mg cm−2 total platinum group metal content) accounts for a quarter of the cost of the PEM cell stack.58 Platinum-free catalysts designed for PEM electrolyzers have made rapid progress in the last decade, and have gained a critical advantage in the potential commercial application of electrolytic technologies.

The leading edge of MoS2 in hydrogen evolution has made it a research hotspot for electrocatalytic HER applications. As a bio-inspired catalyst due to the similarity of its edge site to the metal coordination environment of hydrogenase and nitrogenase enzymes (Fig. 3a),59 MoS2 has been showing attractive hydrogen evolution properties in its two-dimensional (2D) forms, i.e., 2H, 1T, and 3R phases. Experimental and theoretical studies have shown that the intrinsic 2D basal plane of 2H-phase MoS2 (the most stable configuration) is inert during HER catalysis, while defective MoS2, including defect types such as edges, vacancies, doping sites, and strain sites, shows moderate hydrogen adsorption strength and thus affords superior HER performance when employed for catalyzing the hydrogen evolution reaction, especially in acidic media. Similar to that in the 2H phase, the defective MoS2 also shows higher electrochemical activity than the intrinsic structure in the 1T and 3R phases.


image file: d3mh00462g-f3.tif
Fig. 3 (a) Nitrogenase FeMo cofactor with three hydrogen atoms bound at the equatorial μ2S sulfur atoms (left), and hydrogenase active sites with one bound hydrogen atom (right). Reproduced with permission.59 Copyright 2005, American Chemical Society. (b) Volcano plot of the exchange current density versus the DFT-calculated hydrogen free binding energy (ΔGH*). The dotted arrow indicates a ΔGH* value of 0 and the location of the MoS2 edge site is labeled. The inset in the top right corner shows an atomically resolved MoS2 particle (60 Å × 60 Å), where the edge and basal plane are marked with a green dashed box and a blue shaded area, respectively. Reproduced with permission.60 Copyright 2007, AAAS. (c) Calculated free energy versus the HER reaction coordinates as a function of S-vacancy percentage. (d) Schematic of the top (top panel) and side (bottom panel) views of MoS2 with strained S-vacancy on the basal plane. The dashed circles are the S-vacancy percentage. (e) ΔGH* as a function of %x-strain (uniaxial tensile strain) at various %S-vacancies. (f) The Eκ relation of a monolayer MoS2 under tensile strain. Increasing strain in the range of 0–10% results in a narrow band gap. (c–f) Reproduced with permission.19 Copyright 2016, Springer Nature.

2.1. Sulfur vacancies in 2-H MoS2 as HER active sites

With in-depth studies, many types of active sites have been discovered in MoS2 for the HER, including edge sites,59,60 1T-phase metallic structures,61,62 sulfur vacancies in 2H-phase MoS2,19 and so on. For the HER, a general descriptor, the Gibbs free energy change (ΔGH*), has been used to evaluate the activity by theoretical calculations since ΔGH* equal to zero indicates optimal adsorption and desorption processes and thus the highest activity. The edge sites show extraordinary activity due to the near-neutral Gibbs free energy (+0.08 eV, Fig. 3b), indicating a near-optimal H adsorption strength; and the catalytic reaction rate was found to be proportional to the density of edge sites;60 however, it is a challenge to distribute uniformly edge-rich tiny nanoparticles on a catalyst support.63 1T-phase MoS2 is enriched with active sites, but it is metastable and undergoes a phase transition to the 2H phase upon heating/aging.2,61 The vast basal plane in 2H-phase MoS2 was long considered inert until previous work revealed the activation and optimization of the basal plane by creating S-vacancies.19,39,64–66 In the next section, we will discuss the S-vacancy-based defects in 2H-phase MoS2, focusing on the underlying mechanism that unveils the critical role of S-vacancies, the reaction kinetics of defective MoS2 with S-vacancies, the engineering approach for S-vacancies based on 2D MoS2, and the defect-enhanced HER performance.

In our previous research, we have demonstrated that the S-vacancy introduces gap states around the Fermi level, allowing hydrogen to bind directly to the exposed Mo atoms.19 Benefiting from the tuning of the electronic structure, the systematic modulation of ΔGH* can be realized by varying the concentration of S-vacancies ranging from 0% to 25%, while the optimal ΔGH* value lies between 12.5% and 15.62% of S-vacancies (Fig. 3c). On top of S-vacancy engineering, the binding energy is further fine-tuned by applying external tensile strain (see Fig. 3d for diagram) to achieve an optimal ΔGH* of 0 eV under a combination of 3.12% S-vacancies with 8% strain or 12.5% S-vacancies with 1% strain (Fig. 3e). The band structure is also finely tuned in the case of S-vacancies and applied strain, leading to a narrowing of the band gap as well as an increase of gap states around the Fermi level (Fig. 3f). Among various S-vacancies, ∼12.5% S-vacancies with a strain of ∼1.35% resulted in enhanced activity with a reduced overpotential (170 mA at 10 mA cm−2), a lower Tafel slope value (60 mV dec−1), and a better turnover frequency (TOFMo) in acidic media, exceeding the performance of the basal plane of the 1T-phase MoS2 or the edge sites of the 2H-phase MoS2.

To elucidate the HER kinetics of the S-vacancy system, we performed a kinetic study using scanning electrochemical microscopy.67 The substrate generation-tip collection (SG-TC) mode was chosen to determine the kinetics parameters such as apparent rate constant k0, electron-transfer coefficient α, and formal potential E0 (Fig. 4a). Under the assumption of a general one-electron transfer model with Butler–Volmer formalism, COMSOL simulations were performed to calculate the time-dependent H2 (CR) (Fig. 4b) and H+ (CO) distributions before theoretically modeling the LSV curves. The equations, image file: d3mh00462g-t1.tif (where Jv, Kfv, Kbv, and f represent i/nFA, the forward reaction rate constant, the backward reaction rate constant, and F/RT, respectively), establish the connection between the simulated LSV curves and kinetics parameters, which are then extracted by identifying the best fit between the simulated LSV curves and experimental measurements (Fig. 4a). As a result, the unstrained MoS2 containing S-vacancy shows the same E0, α, but a lower k0 than the strained one. Moreover, we verified those kinetic rate parameters for the defective MoS2 electrode through a transient response study, and the results confirmed the reliability of the one-electron reaction model with a Butler–Volmer equation in dealing with the electrocatalytic behavior of MoS2 with S-vacancies or a MoS2-like catalytic system.


image file: d3mh00462g-f4.tif
Fig. 4 (a) Schematic of the scanning electrochemical microscopy setup operating in SG-TC mode (left panel). The top trapezoid represents the Pt ultramicroelectrode (UME) tip where the HOR occurs, and SV-MoS2 at the bottom is the working electrode where the HER occurs. The HOR and HER voltammograms (solid lines) and the corresponding simulation results (dashed lines) for V-MoS2 and SV-MoS2 are compared (right panel). (b) COMSOL simulated spatial distribution of hydrogen concentration (CR/mM). The left side shows the enlarged view of the tip and substrate 2-D configuration. (a and b) Reproduced with permission.67 Copyright 2016, American Chemical Society. (c) Surface energy per unit cell (γ) under applied potentials (URHE) for S-vacancy concentrations varying from 0 to 21.9%. Inset: schematic of stable MoS2 at 12.5% S-vacancies, where the red circle indicates S-vacancies that follow a zigzag configuration. (d) Free energy diagram for protonation and removal of S by an electrochemical (EC) desulfurization method. The green line indicates the free energy path at the applied potential required to make all paths exergonic. A diagram of the EC desulfurization process is shown above, where the indigo, yellow, and cyan balls represent Mo, S, and H atoms, respectively. (e) TOF per Mo atom (TOFMo) versus the applied potential for monolayer MoS2 before (P-MoS2) and after (V-MoS2) EC desulfurization. The performance of V-MoS2 by Ar plasma is also shown in the shaded area. (c–e) Reproduced with permission.65 Copyright 2017, Springer Nature.

There is a wide range of methods for S-vacancy generation in 2H-phase MoS2, including remote hydrogen plasma treatment,68 laser ablation,69 oxygen plasma treatment,70 and helium ion irradiation.71 However, despite the great success in understanding the role of S-vacancies in the HER process, previous methods for the vacancy generation are confined to treatment with argon plasma and similar atmospheres, which act only on the oriented surface of the catalyst due to the anisotropy of plasma and thus have difficulty in penetrating the interior bulk and are not applicable to powder or bulk materials, restricting the production of industrial-scale catalysts. Chemical methods are considered as the alternative solution to generate S-vacancies in any forms of MoS2.

We fabricated S-vacancies in MoS2 by an electrochemical desulfurization process.65 The reduced energy after S removal can be obtained according to theoretical study and the defective sites can be further stabilized with lower surface energy after H occupies the missing S atom, thus preventing the recovery of vacancies or the formation of permanent S-vacancies. As for the vacancy concentration, the surface energy per unit cell (defined as γ) continues to decrease with increasing S-vacancies at a potential of −1 VRHE until a milestone of 12.5% S-vacancies is reached, and the lowest surface energy is obtained for the form of clustered vacancies following a continuous zigzag configuration (Fig. 4c). More than 12.5% of S-vacancies is hard to achieve due to the increased energy cost. Note that the application of a negative potential throughout the electrochemical desulfurization process is extremely necessary. This is further verified by comparing the intermediate steps, where the first protonation (*S + H+ + e → *SH) to form *SH and the subsequent H2S formation step (*SH + H+ + e → H2S + *) overcome huge energy barriers, while both steps become exergonic at a potential of −1.24 VRHE (Fig. 4d). Experimentally, the electrochemical desulfurization method relied on voltage and duration to generate S-vacancies. It was not only applicable to monolayer MoS2, but also polycrystalline multilayer MoS2, obtaining a current increment of about 12 times at −0.32 VRHE. This electrochemical desulfurization was found to be reliable for creating S-vacancies by comparing its electrochemical HER performance with that of other methods (Fig. 4e), indicating that the electrochemical desulfurization approach can effectively replace the argon plasma treatment for creating S-vacancies on the basal plane of the MoS2 electrocatalyst.

Beyond the electrochemical desulfurization route, other feasible chemical approaches have been recently developed with opportunities for large-scale fabrication. From our perspective, three routes are accessible to generate S-vacancies with excellent catalytic properties, including the “reduction route”, “oxidant etching route”, and “doping route”. Specifically, reductants (e.g., lithium, zinc, hydrogen, and NaBH4)40,66,72,73 can react with sulfur to form sulfides under appropriate conditions (e.g., heating) and thus remove sulfur atoms from the basal plane, while oxidants (e.g., H2O2, NaClO)74,75 can also introduce uniformly distributed S-vacancies through a mild etching strategy, and the other doping routes involve the direct chemical bonding between heteroatoms [e.g., palladium (Pd), Ru]76,77 and sulfur atoms to create vacancy sites.

Wu and coworkers proposed a co-annealing method based on a homogeneous mixture of Zn metal and MoS2, in which the formation of a Zn–S bond led to S-vacancies and in situ doping of Zn in the MoS2 lattice, thereby introducing a large number of vacancy defects in the basal plane.40 Theoretical calculations showed that the Zn-doped sites result in a significant decrease in the formation energy of S-vacancies, making it easier to form S-vacancies near the Zn sites and the introduction of S-vacancy induces new states in the band gap near the Fermi level, which might be responsible for the enhanced H atom adsorption at the S-vacancies. This strategy led to a reduction of the HER overpotential from 540 mV of pristine MoS2 to only 194 mV of the defective MoS2 at a current density of 10 mA cm−2, while the Tafel slope value was reduced from 199 to 73 mV dec−1 in an acidic environment. Wang and colleagues developed a simple and gentle H2O2 chemical etching strategy by immersing pristine MoS2 in a H2O2 solution to obtain a homogeneous distribution of single S-vacancies onto the MoS2 nanosheet surface, as captured by aberration-corrected scanning transmission electron microscopy (Fig. 5a).74 Furthermore, by systematically adjusting the etching duration, etching temperature, and etching solution concentration, the comprehensive tuning of S-vacancies can be achieved to obtain the optimal HER performance with a Tafel slope of 48 mV dec−1 and an overpotential of 131 mV at a current density of 10 mA cm−2 in acidic media. The superior activity is attributed to more efficient tuning of the surface electronic structure, which leads to a higher electron density around Mo atoms and a deficiency of electrons on S-vacancy sites (Fig. 5b). The delocalized electrons around the Mo atom with single-atom S-vacancies show stronger attraction to hydrogen atoms, and thus favor the adsorption of hydrogen. Wang and coworkers fabricated a single atom Ru-doped 2H-phase MoS2 catalyst using a two-step hydrothermal method, where Ru replaced some Mo atoms and created S-vacancies, delivering a lower Gibbs free energy of H-adsorption theoretically and a higher activity with an overpotential of 168 mV at a current density of 10 mA cm−2 in 0.5 M H2SO4 experimentally.76 Certainly, not limited to these three current routes, more low-cost and easy-to-operate chemical methods need to be further explored to achieve large-scale fabrication of MoS2 catalysts with S-vacancy defects, which will help in the screening of industrially applicable materials.


image file: d3mh00462g-f5.tif
Fig. 5 (a) STEM image of a CVD-grown monolayer MoS2 flake after etching. The S-vacancy is indicated by yellow dotted circles. (b) Top-view (top) and side-view (bottom) electron density difference maps for pristine (left panel), defective MoS2 with agglomerated S-vacancies (middle panel) and single S-vacancies (right panel). The orange and blue balls represent S and Mo atoms, respectively. (a and b) Reproduced with permission.74 Copyright 2020, American Chemical Society. (c) Proposed engineering procedures to create defective 2H-phase MoS2 for the HER, which include phase I (materials synthesis), phase II (materials engineering, e.g., N-doping, hybrid, vacancy, strain and their combinations), and phase III (electrochemical H2 production). (d) Experimental TOFS-vacancy values versus their corresponding calculated ΔGH* values for MoS2 with S-vacancies before applying strain (V-MoS2) and after applying strain (SV-MoS2). The dashed line indicates a volcano relation. (e) Overpotential of Co/MoS2−x and MoS2−x at −10 mA cm−2 (left, y-axis) and the loading of electrodeposited Co on Co/MoS2−x (right, y-axis) versus desulfurization voltage. Reproduced with permission.79 Copyright 2018, American Chemical Society. (f) Colored contour plot of surface energy per unit cell γ (with respect to the bulk MoS2) as a function of S-vacancies and uniaxial strain. (d and f) Reproduced with permission.19 Copyright 2016, Springer Nature. (g) The adhesion energy per carbon atom of pristine MoS2 on defective graphene with carbon vacancy%. Insets: Atomic configurations of MoS2 on various defective graphene substrates with a graphene layer on top. The blue, yellow, and grey balls represent Mo, S, and C atoms, respectively. Cyan balls indicate the missing carbon atoms. (h) The adhesion energy per carbon atom of MoS2 with S-vacancies on pristine graphene. Insets: Atomic configurations of MoS2 with various S-vacancy% on a graphene substrate with MoS2 layer on top. The red balls indicate the missing sulfur atoms arranged in a line shape. (g and h) Reproduced with permission.80 Copyright 2020, Elsevier. (i) The reaction energy versus reaction path for Pd doping into MoS2. Insets: Simulated atomic structures, where the blue, pink, and yellow balls represent Mo, Pd and S atoms, respectively. Path 1 (the black pathway) is thermodynamically more favorable due to its lower defect formation energy. Reproduced with permission.81 Copyright 2020, Elsevier.

Although important breakthroughs have been made for deepening the understanding of the mechanism of S-vacancy-enhanced activity, for opening new defects types (e.g., Frenkel defects caused by S-vacancies in monolayer MoS2),78 for observing the kinetics behavior of the defective catalyst, and for developing advanced fabrication methods of S-vacancies in single-crystalline monolayers or polycrystalline multilayers, a crucial issue to be addressed is the further improvement of activity and stability of MoS2 with S-vacancies in acidic media, which is critical for practical applications. The inherent structure and properties determine the upper limit of the catalytic activity of MoS2 with S-vacancies, and the breakthrough in performance can be achieved by combining S-vacancies with other engineering methods.

It is worth noting that several engineering approaches, including but not limited to N-doping, hybridization, vacancy engineering, and strain engineering, can be integrated into a single MoS2 system for more efficient HER catalysis (Fig. 5c). We have applied tensile strain to MoS2 with S-vacancies to tune the binding energy and band structures, which leads to an increase in catalytic activity, as evidenced by lower overpotentials and higher TOF values (see Fig. 5d for the TOFs) in an acidic environment.19 The catalytic properties of S-vacancies can also be tuned by integrating with heteroatoms. Kang's DFT calculations showed that by doping the transition metals, such as Sn, Tc, Ir, Rh, Ru, Re, Os, Pd, and Pt, into MoS2 with S-vacancies and enabling the doped atom to replace the Mo atom adjacent to the S-vacancies, the HER activity of defective MoS2 can be improved.82 Park and coworkers loaded Co clusters onto S-vacancies to form a Co–Mo preferentially active interface,79 and the calculated ΔGH* value for Con–MoS2−x is closer to the optimal value of zero. Experimentally, the Co addition on the vacancy enabled the desulfurized MoS2−x multilayer catalyst to attain a better activity with optimized overpotentials at a current density of −10 mA cm−2 in 0.5 M H2SO4 electrolyte (Fig. 5e). Similarly, the interactions between S-vacancies and other processes, such as Ru deposition,83 Zn doping,40 Pd doping,81 and O substitution84 have all shown enhanced HER performance. Table 3 lists more HER catalysts built on 2H-phase MoS2 with S-vacancies in acidic media. Moreover, the combination of S-vacancies and phase modulation is also useful to elevate the catalytic activity, which will be discussed later.

Table 3 Summary of HER performance of 2H-phase MoS2 catalysts with S-vacancies in acid
CatalystRef. Methods for S-vacancies H adsorption free energy Density of S-vacancies Overpotential (vs. RHE) Tafel slope (mV dec−1) TOF Stability
a The symbol “∼” represents estimated values from the figures in the reference articles.
Monolayer of strained MoS2 with S-vacancies19 Ar-plasma treatment 0 eV at 12.5% S-vacancies and 1% strain 12.5 ± 2.5% with strain 1.35 ± 0.15% 170 mV@10 mA cm−2 60 TOFS-vacancy 0.31 s−1@0 V N/A
S-vacancies and zinc doping on MoS240 Zinc reduction −0.08 eV 13% 194 mV@10 mA cm−2 78 N/A 1000 cycles
S-vacancies of MoS264 CVD process −0.095 eV 7–10% ∼200 mV@10 mA cm−2 65–75 TOFS-vacancy 3.2 ± 0.4 s−1@0 V 10[thin space (1/6-em)]000 cycles
S-vacancies on 2H-phase MoS265 Electrochemical desulfurization Within 0.1 eV (15 ± 8)% ∼9.5 mA cm−2@320 mVa 193 (Multilayer) TOFMo 2 s−1@0.135 V 8 h@−0.32 V
MoS2 with ultrarich S-vacancies66 Annealing under hydrogen N/A up to 90% ∼320 mV@10 mA cm−2 ∼83 TOFsurface[thin space (1/6-em)]S 2 s−1@η ≈ 0.235 V 24 h@10 mA cm−2
S-vacancies in the basal plane of monolayer MoS268 Remote hydrogen plasma ∼−0.06 eV 24% 183 mV@10 mA cm−2 77.6 TOFS-vacancy ∼1.42 s−1@0 V N/A
S-vacancies in the basal plane of 2H-phase MoS269 Laser ablation −0.17 eV ∼8% 178 mV@10 mA cm−2 41.4 7.674 s−1@η = 0.25 V 10 h@−0.178 V
Defective MoS2 with S-vacancies70 Oxygen plasma exposure N/A N/A 620 mV@10 mA cm−2 171 N/A 10[thin space (1/6-em)]000 cycles
Single layer MoS2 with vacancies71 Helium ion irradiation N/A 5.7 × 1014 cm−2 ∼200 mV@100 mA cm−2 44 TOFsingle[thin space (1/6-em)]atomic vacancy 5 s−1@0 V N/A
S-vacancies in the basal plane of MoS272 Electrochemical lithium-ion intercalation 0.70 eV ∼18% 200 mV@10 mA cm−2 65 0.0931 s−1@η = 0.25 V 80[thin space (1/6-em)]000 s @-200 mV
S-vacancies in MoS2 basal plane73 NaBH4 reduction N/A ∼12% 190 mV@10 mA cm−2 54 0.38 s−1@−0.2 V 15 h@10 mA cm−2
Single S-vacancies onto MoS274 H2O2 chemical etching 0.02 eV 12.11% 131 mV@10 mA cm−2 48 N/A 12 h@10 mA cm−2
Ru-doping 2H-phase MoS2 with S-vacancies76 Two-step hydrothermal method −0.16 eV N/A 167 mV@10 mA cm−2 77.5 0.23 s−1@−0.1 V 12 h@10 mA cm−2
Pd-MoS2 with S-vacancies77 Spontaneous interfacial redox reaction −0.02 eV 16.7% 78 mV@10 mA cm−2 62 TOFS-vacancy+Pd−S* 0.15 s−1@0 V 100 h@10 mA cm−2
Frenkel-defected monolayer MoS278 Annealing in Ar 0.36 eV 0.85% 164 mV@10 mA cm−2 36 N/A 10 min@10 mA cm−2
Co/MoS2−x79 Electrochemical desulfurization and electrodeposition Within 0.1 eV N/A 201 mV@10 mA cm−2 82 N/A 1000 cycles
MoS2 with line-shaped S-vacancies on VGN80 Ar plasma exposure <0.16 eV 12% 128 mV@10 mA cm−2 50 N/A 500 h@10 mA cm−2
Pd doping-induced S-vacancies in MoS281 Solvothermal and substitution reactions −0.01 eV N/A 106 mV@10 mA cm−2 60 N/A 1100 h@10 mA cm−2
2D MoS2−xOx solid solution crystal84 Oxygen substitution during exposure 1.2 eV 5 × 1013–1 × 1014 cm−2 ∼260 mV@10 mA cm−2 67 N/A 1000 cycles


Nevertheless, S-vacancies disrupt the crystalline structure of MoS2, thus increasing the S-vacancy ratios increases the surface energy per unit cell (γ) (Fig. 4c and 5f),19,65 which makes MoS2 less stable. It is difficult to maintain the catalytic performance over long periods due to the structural instability caused by S-vacancies. Therefore, several strategies have been proposed to achieve high stability by stabilizing S-vacancies on defective MoS2 catalysts, such as lowering the systematic energies by means of a stable support,80 anchoring a heteroatom such as Pt or P at the S-vacancy sites,85 and theoretically stabilizing the S-vacancy sites by Jahn–Teller distortion.86 Moreover, the higher activity with smaller polarization also enhances the durability as the catalyst operating at a smaller overpotential can reduce the damage to the structure, although for the practical catalyst there is often a trade-off between activity and stability, e.g., the introduction of support materials or heteroatoms could block the highly active S-vacancies.

Recently, we reported a catalyst that stabilizes S-vacancies of 2H-phase MoS2 supported on a vertical graphene network.80 The introduced S-vacancies into MoS2 and C vacancies into graphene synergistically modulate the stability and activity. The DFT results showed that adhesion of MoS2–graphene systems is exergonic and thus these systems are relatively stable. Among the systems with or without defects, the weakest adhesion system is pristine MoS2 over pristine graphene, while C vacancies or energetically favorable line-shaped S-vacancies drastically improve the adhesion energy between the catalyst and support (Fig. 5g and h). Specifically, 5.56% C vacancies and 12.50% S-vacancies achieve an 8.15% and 12.49% increase in adhesion energies, respectively. The catalyst with 12% S-vacancies showed the optimal activity with a 128 mV of overpotential at 10 mA cm−2 and achieved stable performance over 500 h in an acidic environment. We also fabricated an ultra-stable electrocatalyst based on MoS2 and graphene by doping with Pd.81 Theoretical studies revealed the preferred pathways for substitutional doping, i.e., formation of a Mo vacancy followed by doping of Pd on a Mo vacancy and then the creation of S-vacancies (Fig. 5i, path 1). Thus, the Mo vacancy sites, grains, boundaries, defects, or the edge of the MoS2 lattice are preferred to be doped firstly. Pd doping generates rich S-vacancies, which increases the H adsorption and thus improves the activity and further increases the interlayer adhesion. The catalyst exhibited an ultra-stable hydrogen evolution operation for up to 1000 h under 10 mA cm−2 with a much lower degradation rate of 23 μV h−1 and for up to 180 h at a high current density of 80 mA cm−2, outperforming many HER catalysts operating in acidic media. It therefore offers an efficient method for making stable, highly active and low-cost catalysts with great potential to replace Pt in PEM electrolyzers.

2.2. Defects in 1-T and 3-R MoS2 as HER active sites

1T-phase MoS2 with octahedral coordination and an expanded interlayer distance possesses metallic electrical conductivity and superior catalytic activity at both basal plane and edge sites,87,88 but also has a metastable structure. Engineering of 1T-phase MoS2 by basic size control, vacancy formation, or grain boundary modulation can enhance HER activity. For example, by synthesizing nanodots to provide high-density active edge sites,89,90 by fabricating porous metallic 1T-phase MoS2 nanosheets with S-vacancies,39,72 or by making 2H-1T crystal boundaries to optimize the adsorption strength of the hydrogen atom.91 Besides, at least three defect types, including structural distortion,92 hetero-element doping,93 and complex engineering,47 have been proposed as efficient activation strategies for HER catalysis. In the next section, we will first discuss the HER application of a distorted 1T-phase MoS2, focusing on its growth strategy and its integration with other defects; then discuss other defect engineering strategies of 1T-phase MoS2, with emphasis on the construction of defects by in situ or non-in situ doping of metals or nonmetals to promote the HER, followed by the identification of dynamic active sites in doped 1T-phase MoS2, and finally we present the HER study of defective 3R-phase MoS2.

A distorted 1T-phase MoS2, referred to as 1T′-phase MoS2, features two adjacent Mo atoms with shortened distances, which are connected to form a zigzag chain, while S atoms in the distorted octahedron are structurally deviated from the pristine plane (Fig. 6a), belonging to the point group C2v and space group Pmn21 in the orthorhombic crystal system.94 MoS2 in the 1T′ phase retains its metallic behavior as in the case of 1T-phase MoS2, benefiting from a similar splitting of the d orbitals into two generated states, while electrons populate the lower energy levels and show a similar distribution of density of states that hosts the Fermi level.7 The subtle difference between 1T-phase MoS2 and 1T′-phase MoS2 is that the structural distortion of the metal chain period doubling leads to an intrinsic band inversion between chalcogenide-p and the metal-d bands around Γ, as well as the spin–orbit coupling opening a fundamental gap of 0.08 eV at the Dirac point (Fig. 6b).95 DFT calculations suggest that the 1T′ phase can be converted spontaneously from the 1T phase without barrier ascribing to its relatively low energy, resulting in improved structural stability of MoS2.96 Controllable fabrication of 1T′-phase MoS2 is important for catalysis, superconductivity, magnetism, and other fields. A common method for producing 1T′-phase MoS2 involves intercalation with alkali metals (e.g., Li, Na, and K ions).7 Nevertheless, this method produces either incomplete phase transition or excessive intercalation, leading to impurity or decomposition of MoS2.


image file: d3mh00462g-f6.tif
Fig. 6 (a) Schematic illustration of 1T′-phase MoS2 structure, where the red dashed line indicates the zigzag chain of Mo atoms along the a-axis. Reproduced with permission.98 Copyright 2019, John Wiley & Sons, Inc. (b) Band structure of 1T′-phase MoS2. Eg and 2δ represent the fundamental gap and inverted gap, respectively. The inset compares band structures with (red dashed line) and without (black solid line) spin–orbit coupling. The letters p and d mainly indicate the chalcogenide p-orbitals and metal d-orbitals. Reproduced with permission.95 Copyright 2014, AAAS. (c) Electrochemical setup for HER measurements on electrochemical microcells, where MoS2 nanosheet, Pt wire, and Ag/AgCl are used as working, counter, and reference electrodes, respectively. (d) Polarization curves of EM-1, EM-2 and EM-3 in 0.5 M H2SO4. Inset: Optical microscopy image of EM-1. The scale bar is 20 μm. (e) Schematic of the fabrication of three types of electrochemical microcells, EM-1 (1T′ phase), EM-2 (1T′ and 2H phases) and EM-3 (2H phase). (c–e) Reproduced with permission.97 Copyright 2018, Springer Nature. (f) Schematic of the phase-controlled synthesis strategy for 1T′ and 2H. The bottom inset shows the calculated formation-energy difference between 1T′-phase KxMoS2 and 2H-phase KxMoS2versus K concentration. (g) Phase diagram of MoS2 grown using a K2MoS4 precursor as a function of growth temperature and H2 concentration. (h) Tafel plots of 1T′- and 2H-phase MoS2 flakes grown by CVD using K2MoS4 precursor. (f–h) Reproduced with permission.94 Copyright 2018, Springer Nature.

More advanced preparation methods have been pursued before exploring the catalytic properties of 1T′-phase MoS2. Peng, Yu, Nam and their colleagues have all recently developed a method to produce the high-purity 1T′-phase MoS2 (as evaluated using the Mo 3d XPS spectra for the percentage of the 1T′ phase in the crystalline phases) using K2MoO4 and S as precursors and multi-annealing processes prior to post-treatment with I2 acetonitrile solution.7,97,98 This method endows MoS2 with high-purity 1T′ phase in the form of bulk, or an exfoliated monolayer with a phase purity of 97% and an unprecedented lateral size up to tens to hundreds of micrometers. As observed, typical zigzag chains appear along the a-axis and two different interlayer distances alternate along the b-axis (Fig. 6a), and the lattice constant of the a-axis is shorter than that of the b-axis due to the structural distortion. In addition to the structural uniqueness, the 1T′-phase MoS2 can also be distinguished by electronic or phonon information. The binding energy of the Mo 3d peak on the spectrum of X-ray photoelectron spectroscopy (XPS) is lower (∼0.8 eV negative shift) compared to that of the 2H-phase MoS2, and only J1, J2, J3 and A1g peaks are present on the Raman spectrum without some characteristic peaks of the 2H (E2g) and 1T (Eg, Ag) phases due to the broken structural symmetry, which both contribute to the identification of the 1T′ form.97,98 In a typical three-electrode measurement (Fig. 6c), the pure 1T′ phase exhibited the best activity with an onset overpotential of 65 mV and a Tafel slope of 100 mV dec−1 (Fig. 6d), which is better than those of the 1T′-2H mixture (128 mV dec−1, obtained by laser-induced transformation as depicted in Fig. 6e) and the 2H phase (180 mV dec−1, obtained by thermal annealing as shown in Fig. 6e) in an acidic electrolyte.97 The superior activity is attributed to the inherent active basal plane and metallic charge transport in the 1T′ phase. Moreover, in-plane anisotropic charge transport affects the catalytic performance along different crystal orientations of the 1T′ phase, as revealed by electrocatalytic evaluation based on an eight-terminal device with four pairs of diagonal electrodes, where the conductivity relies on angle-resolved crystal orientation and the better charge transport ability leads to slightly enhanced catalytic activity.98

Towards the goal of achieving an even higher purity of 1T′-phase MoS2, Liu and coworkers proposed a one-step CVD growth route using K2MoS4 as a precursor to realize approximately 100% phase purity in the monolayer form.94 Theoretical study suggests that potassium ion reduces the formation energy of the 1T′ phase theoretically and the K+ concentration exceeding 44% renders the 1T′ phase with a more stable structure than the 2H-phase MoS2 (Fig. 6f). The synthesis of 1T′-phase MoS2 undergoes the reaction process, K2MoS4 + H2 → KxMoS2 + K + K2S + H2S (gas), where the annealing atmosphere (5–12% H2 concentration in argon), temperature (650–750 °C) (Fig. 6g), and the substrate (fluorophlogopite mica) play crucial roles in the production of the target 1T′ monolayer material. The K+ concentration in the product was controlled by tuning the ratio between reductive H2 and inert argon gases in the experiment, and a 50% K+ concentration was observed to stabilize the CVD-grown 1T′-phase MoS2 flakes instead of the 2H phase. K+ ions can be completely removed after rinsing with water prior to evaluating the properties. Metallic 1T′-phase MoS2 shows better HER performance than the 2H-phase flakes with an onset potential of 205 mV and a Tafel slope of 51 mV dec−1 (Fig. 6h), attributed to its faster electron transport and higher density of active sites. Notably, the current density exhibits negligible loss after 1000 cycles, demonstrating its decent stability in an acidic environment.

Liu and colleagues further created the defects based on the 1T′-phase MoS2.92 A bottom-up colloidal strategy was developed to prepare high-phase-purity, nano-monolayer, and strained 1T′-phase MoS2 with Mo vacancies. The 1T′-phase MoS2 nano-monolayer was subjected to ∼3% tensile strain in the xx direction, ∼2% shear strain in the xy direction (Fig. 7a), and 2% compressive strain in the yy direction according to the geometric phase analysis (GPA) using the atomic-scale HRTEM image. The Mo vacancies and strain together tune the band structure and hydrogen adsorption free energy (Fig. 7b). The best HER performance was achieved on the active ST site at 3% uniaxial tensile strain. Also, the abundant in-plane/edge active sites and high electrical conductivity contribute to its electrocatalytic properties. In the acidic HER measurement, an overpotential of 149 mV at a current density of 10 mA cm−2 and a Tafel slope value of 42 mV dec−1 were obtained, with the durability for up to 48 h at an overpotential of 165 mV.


image file: d3mh00462g-f7.tif
Fig. 7 (a) Strain map of the 1T′-MoS2 nano-monolayer. (b) Theoretically calculated free energy of H adsorption at different S sites under different strain conditions for the 1T′-phase MoS2 structure with a Mo vacancy. (a-b) Reproduced with permission.92 Copyright 2022, American Chemical Society. (c) Wavelet transform (WT) contour plots of MoS2/VG (top) and (N,PO43−)-MoS2/VG (bottom). VG: vertical graphene. The R values of 1.93 Å, 2.92 Å, and 1.45 Å are associated with Mo–S, Mo–Mo and Mo–N bonding, respectively. (d) Top panel: the diagrams of bonding and antibonding of 2H-phase MoS2 and (N, PO43−)–MoS2. Bottom panel: The density of states (DOS) plots of 2H-phase MoS2, PO43−–MoS2, N–MoS2 and (N, PO43−)-MoS2 and the corresponding d-band center potentials. (c and d) Reproduced with permission.47 Copyright 2019, John Wiley & Sons, Inc. (e) Hydrogen adsorption free energy on the basal plane of pristine 1T-phase MoS2 and intercalated 1T-phase MoS2 with Ca, Na, Ni, and Co. Reproduced with permission.99 Copyright 2017, American Chemical Society. (f) HAADF-STEM of Pt-1T-SMoS2 (left panel) and ADF intensity line profiles (right panel) taken along the corresponding color lines (black, blue, and red). Reproduced with permission.100 Copyright 2019, American Chemical Society.

Apart from 1T′ phase engineering, doping of 1T-phase MoS2 is another effective strategy to enhance the electrocatalytic performance. Tang and coworkers used DFT simulations to screen eleven transition metal dopants and found that Mn, Cr, Cu, Ni, and Fe can reduce the ΔGH* value of the doped 1T phase.101 Experimentally, at least three doping methods have been developed, including one-step doping synthesis, post-synthesis doping, and doping-induced phase transition. Regarding the one-step doping method, Ji and coworkers pointed out that Cu atoms can be doped into MoS2 by reducing MoS3 with intermediate Cu+ species under hydrothermal conditions to form Cu-bound metal 1T-phase MoS2, where Cu can be inserted into the S layer and form Cu–S bonds.93 The synergistic effect of 1T-phase MoS2, doped Cu atoms, and S- vacancies contributes to its elevated activity with a relatively low overpotential (131 mV@10 mA cm−2) and a Tafel slope of 51 mV dec−1. Similar to the function of Cu, Jin and coworkers found that the introduced Co atoms also contribute to the HER activity by expanding the interlayer space, enriching the Co–Mo–S active sites (40.9–91.3%) and the doped 1T phase (73.9–79.2%).102 Specifically, defective MoS2 was prepared using a hydrothermal approach with high-pressure H2 utilized for expanding the interlayer space, while Co atoms form Co–Mo–S in the product. Polarization curves indicated that the as-prepared catalyst displayed an outstanding overpotential of 59 mV at 10 mA cm−2 and a Tafel slope of 32 mV dec−1 in acidic media, implying the obvious function of Co atom doping and H2 modulation on the defective 1T-phase MoS2 system.

Beyond metal atom doping, doping with nonmetallic elements into 1T-phase MoS2 is also effective. Deng and coworkers recently developed a synergistic N-doping plus PO43− intercalation strategy to realize the enlarged interlayer distance, high 1T phase proportion (41%), and nitrogen doping into the lattice with the introduction of the Mo–N bond (Fig. 7c).47 The d-band center of (N,PO43−)-MoS2 is lower than that of 2H-phase MoS2 (Fig. 7d), leading to more electrons filling the antibonding states and thus weakening the interaction between the valence states of the adsorbent and catalyst. (N,PO43−)-MoS2 achieves an optimized hydrogen adsorption free energy (ΔGH* = 0.07 eV) for the HER and the combination of (N,PO43−)–MoS2 with VG skeleton further enhances the electrical conductivity. Together with these advantages, the (N,PO43−)–MoS2/VG catalyst exhibited superior HER activity with a small overpotential (85 mV@10 mA cm−2) and a low Tafel slope (42 mV dec−1), as well as decent stability for 10 h at a current density of 10 mA cm−2 in acidic media. Tan and colleagues synthesized 1T-phase (71%) MoSSe single-layer nanodots with Se vacancies by a combination of ball milling and chemical Li-intercalation.90 The alloying effect (i.e., alternative H–S or H–Se bond) and Se vacancies (in this case, H adsorbs on the periphery S around Se vacancies) enable MoSSe nanodots to acquire a more neutral ΔGH*. Hence, the 1T-phase MoSSe nanodots yielded a low overpotential of 140 mV at 10 mA cm−2 and a Tafel slope of 40 mV dec−1, as well as superior durability for 10[thin space (1/6-em)]000 cycling in an acidic electrolyte.

A post-synthesis doping procedure is performed on the as-synthesized 1T-phase MoS2 by a gentle process that avoids phase destruction. Attanayake and coworkers reported a precipitation route in which solutions containing Na+, Ca2+, Co2+, or Ni2+ ions are added into the 1T-phase MoS2 nanosheets, which results in 4–5 atomic% intercalation without altering the 1T phase as well as the 1T proportion (∼75%), leading to optimized H adsorption free energy ΔGH* (Fig. 7e) and elevated HER performance compared to the pristine 1T phase.99 Moreover, Lau and coworkers pointed out that sonochemical doping, a simple modulation method under ambient conditions to prevent the reconstruction from the 1T phase to the 2H phase, can achieve the inclusion of transition metal atoms, such as Pt or Pd, at the surface/interface of 1T-phase MoS2.100 The sonochemical process with a reduction function towards metal precursors led to the chemisorption of single transition metal atoms on the basal plane of exfoliated 1T-phase MoS2, as captured by high-angle annular dark-field imaging-scanning tunneling electron microscopy (HAADF-STEM) images (Fig. 7f). It was significantly more active than the pristine 1T phase after doping, reaching an overpotential of 140 mV (3 wt% Pd doping) and 223 mV (3 wt% Pt doping) at 10 mA cm−2 in 0.5 M H2SO4 electrolyte.

Single-atom doped 1T-phase MoS2 can also be achieved by atom-induced phase transition. Qi and coworkers proposed the formation of the 1T phase induced by cobalt nanodisk addition and subsequent doping with cobalt by sonication and leaching methods (Fig. 8a).103 A cobalt nanodisk was found to efficiently trigger the phase transition to distorted 1T-phase MoS2 (D-1T MoS2, see Fig. 8b), arising from the strain caused by the lattice mismatch between metallic Co and 2H-phase MoS2, while the Co–S bonds are also formed with the assistance of sonication (Fig. 8c). The disordered structure in MoS2 emerged upon the incorporation of Co, which is evenly located on top of the Mo atom in a single-atom site and coordinated to the adjacent three sulfur atoms, while higher Co coverage (Co–Co distance less than 3.10 Å with a strain of ∼3.70%) determines the stable presence of the 1T phase, as suggested by DFT. SA Co-D-1T MoS2 with 3.70% cobalt coverage shows the lowest ΔGH* value of 0.03 eV, very close to the ideal 0 eV. In the actual experiment, 3.54% Co loading was introduced into D-1T MoS2, delivering an excellent HER performance with a 42 mV overpotential at 10 mA cm−2, a 32 mV dec−1 Tafel slope value that is superior to most nonprecious catalysts (Fig. 8d), and robust stability for over 10 days (at η = 100 mV vs. RHE) in acidic media.


image file: d3mh00462g-f8.tif
Fig. 8 (a) Schematic diagram of the fabrication process for an atomic cobalt array covalently bound to distorted 1T-phase MoS2 nanosheets (SA Co-D 1T MoS2). (b) Aberration-corrected HAADF-STEM image of SA Co-D 1T MoS2, showing the distinct junction between SA Co-D 1T MoS2 and pristine 2H-phase MoS2. The inset shows the EELS spectrum with peaks at 779 eV and 794 eV, corresponding to Co L3 and Co L2, respectively. The scale bar is 1 nm. (c) FT-EXAFS spectra of SA Co-D 1T MoS2, Co NDs-MoS2, and cobalt foil at the Co K-edge. A single strong shell in R-space indicates the exclusive presence of the Co–S bond for SA Co-D 1T MoS2. (d) Comparison of HER activity using Tafel slope versus overpotential at a current density of 10 mA cm−2. A decrease in the Tafel slope or overpotential indicates better activity. Pt/C and SA Co-D 1T MoS2 are marked in the dashed box. (a–d) Reproduced with permission.103 Copyright 2019, Springer Nature. The R-space FT-EXAFS spectra (open points) and the fitting results (solid lines) of Ni@1T-MoS2 at (e) Mo K-edge and (f) Ni K-edge via in situ XAS. (g) Proposed active species of Ni@1T-MoS2 under acidic conditions at −0.76 VRHE. Ni remains predominantly coordinated by S atoms and is reduced at the applied potential. (e–g) Reproduced with permission.104 Copyright 2020, Springer Nature.

Despite considerable progress in exploring element selection, synthetic methods, electrochemical performance, and DFT theoretical understanding, tracking and identifying the dynamic active sites of defective 1T-phase MoS2 during HER operation remain ambiguous despite its essentiality for understanding the HER. Pattengale and coworkers recently revealed the dynamic behavior of active sites by capturing the intermediate structure of a doped 1T-phase MoS2 electrocatalyst.104 In their Ni@1T-MoS2 catalytic system (Fig. 8e and f), a shortening of the Mo–S distance at an applied potential of −0.76 V under acidic conditions was observed, while the intrinsic Ni site did not undergo significant changes in the first coordination shell despite a shortening of the Ni–Mo second-shell interaction. Single-atom Ni replacing Mo on the basal edge underwent reduction from the Ni2+ oxidation state to lower valence state at externally applied potentials, without significant change of the pristine structure where the single Ni site bonded to surrounding Mo, S, and O (Fig. 8g). Thus, 1T-phase MoS2 as the active phase and Ni as the active species undergo minimal changes when no potential, 0 V, and −0.76 V are applied, indicating that the intrinsic catalytic structure of Ni@1T-MoS2 is quite stable during the dynamic evolution of the catalyst under acidic conditions. Inspired by this work, similar work is proposed to identify the active sites of defective MoS2 during dynamic catalysis to deepen the understanding of the MoS2-based catalyst systems. Besides, the controllability including defect concentration, defect location, electrical conductivity, and the proportion of 1T phase in defective MoS2 remains challenging and needs more attention in future studies.

In stark contrast to the defective 1T phase, the semiconductor 3R-phase MoS2 with R3m space group and a ABC–ABC stacking order structure possesses unique properties such as excellent non-linear optical response, valley-dependent spin polarization benefiting from broken inversion symmetry even in the bulk form, offering huge opportunities for applications in nonlinear optical devices, electronic and optoelectronic devices, etc.105,106 In the field of HER catalysis, Toh and coworkers reported that a mixed-phase MoS2 consisting of rhombohedral 3R and the hexagonal 2H (in a ratio of 2.3[thin space (1/6-em)]:[thin space (1/6-em)]1) phases obtained by a flux route with MoO3 and S as the precursors in carbonate media at 550 °C, shows a Tafel slope of 113 mV dec−1 (see Table 4 for the comparison of the Tafel slopes and other parameters for more defective 1T- and 3R-phase MoS2), superior to that of pure 2H-phase MoS2 (170 mV dec−1) and comparable to that of the exfoliated 1T-phase MoS2 (99 mV dec−1) under acidic conditions.107 However, more defective 3R-phase MoS2 with enhanced catalytic properties remains underexplored, although the S-vacancy defects have also been uncovered in the 3R phase and these defects exhibit migration phenomena similar to that of the 2H phase.108 Therefore, the methods for making vacancy, strain, doping, and other defects that could hold promise in enhancing the catalytic performance of 3R-phase MoS2 need further development.

Table 4 Summary of HER performance of defective 1-T and 3-R MoS2 catalysts in acid
CatalystRef. Method for defects H adsorption free energy 1T(T′)/3R phase fraction Overpotential (vs. RHE) Tafel slope (mV dec−1) Stability Loading (mg cm−2)
Porous 1T-phase MoS2 nanosheets39 Liquid-ammonia-assisted lithiation N/A ∼82% 1T phase 154 mV@10 mA cm−2 43 15% decay after 20[thin space (1/6-em)]000 s@η = 180 mV 0.14 ± 0.01
(N,PO43−)-MoS2/VG47 Hydrothermal and annealing 0.07 eV 41% of 1T phase 85 mV@10 mA cm−2 42 10 h@10 mA cm−2 N/A
S-intercalated 1T′-phase MoS2 anchored on GNRs87 Solvothermal process 0.06 eV 1T′[thin space (1/6-em)]:[thin space (1/6-em)]2H = 1[thin space (1/6-em)]:[thin space (1/6-em)]1.3 205 mV@10 mA cm−2 50 24 h@10 mA cm−2 0.375
Metallic-phase MoS288 Hydrothermal process N/A ∼86% of metallic-MoS2 175 mV@10 mA cm−2 41 <12% decay after 1000 cycles ∼0.043
2D-MoS2 QDs89 Quasi-full electrochemical lithiation N/A 94% of 1T phase 92 mV@10 mA cm−2 44 80 h@200 mA cm−2 0.1
MoSSe nanodots90 Ball milling and Li-intercalation 0.19 eV 71% of 1T phase 140 mV@10 mA cm−2 40 Negligible change after 10[thin space (1/6-em)]000 cycling 0.071
MoS2 with 2H-1T boundaries91 Ar-plasma induces the 2H-to-1T-phase −0.13 eV N/A 136 mV@10 mA cm−2 73 200 h@η = 150 mV N/A
1T′-phase MoS2 nano-monolayers92 Bottom-up colloidal synthesis <0.05 eV N/A 149 mV@10 mA cm−2 42 48 h@20 mA cm−2 0.07/0.2
Porous 1T-phase MoS2 with Cu doping93 One-pot solvothermal method N/A N/A 131 mV@10 mA cm−2 51 25[thin space (1/6-em)]000 s@10 mA cm−2 0.1
1T′-phase MoS2 flakes94 K-assisted CVD with K2MoS4 precursor N/A Approximately 100% 1T′ phase purity 205 mV@10 mA cm−2 51 <5% loss of after 1000 cycles N/A
1T′-phase MoS2 nanosheet97 Annealing of K2MoO4 and S and exfoliation N/A 91% of 1T′ phase 175 mV@10 mA cm−2 100 N/A N/A
Na/1T-MoS299 Lithium intercalation and Na+ addition <0.05 eV 75% of 1T phase 183 mV@10 mA cm−2 45 24 h@10 mA cm−2 0.71
Pd-1T-SMoS2100 Sonochemical-assisted reductive doping 0.17 eV 76% of 1T phase 140 mV@10 mA cm−2 50 10 mV increase after 1000 cycles N/A
Co–MoS2-n102 Hydrothermal process 0.09/0.07 eV 76.8% of 1T phase 56 mV@10 mA cm−2 32 12 h@η = 59 mV 0.29
SA Co-D 1T MoS2103 Sonication assembly and CV leaching 0.03 eV N/A 42 mV@10 mA cm−2 32 10 days@η = 100 mV 0.28
Ni@1T-MoS2104 Hydrothermal growth N/A N/A ∼220 mV@10 mA cm−2 N/A 1 h@η = 760 mV 1
3R-phase MoS2107 Thermal reaction N/A 2.3[thin space (1/6-em)]:[thin space (1/6-em)]1 of 3R/2H ratio 520 mV@10 mA cm−2 113 N/A 0.057


It is worth reminding that the defective MoS2 in the 2H/1T/3R phase discussed above is applied to catalyze the HER in acidic media. In contrast, intrinsic MoS2 generally behaves poorly for the HER in an alkaline environment due to the high activation energy of the Volmer step of water dissociation and the strong adsorption of the formed OH on the MoS2 surface.109 Although S-vacancy or mixed-phase defects can also effectively boost the HER kinetics of MoS2, the increased activity from these defects is still far below that of the Fe/Co/Ni-based composition.110,111 Therefore, most of the current studies on MoS2-based catalysts for the HER in alkaline media have focused on the introduction of Ni or Co components in the form of atom/oxide/hydroxide/sulfide/nitride/phosphide109,111–115 and some studies are on the construction of heterojunctions with other compounds116 or on the introduction of noble metals.117

For example, Liu and coworkers employed hydrothermal and sulfidation methods to prepare 1T0.72-MoS2@NiS2, where heterogeneous interfaces of MoS2 and NiS2, heterojunctions of 2H and 1T phases, numerous defects, and high 1T ratios were observed.111 When catalyzing the HER in an alkaline environment, 1T0.72-MoS2@NiS2 exhibited an overpotential of 128 mV and stability for 30 h at a current density of 10 mA cm−2. Moreover, it had a Tafel slope value of 68 mV dec−1, which is lower than those of 1T0.41-MoS2 (89 mV dec−1), 2H-phase MoS2 (124 mV dec−1), and 1T-phase MoS2 (198 mV dec−1). DFT calculations showed that 1T-MoS2@NiS2 has a metallic phase with zero band gap, and PDOS (projected density of states) intensity of 1T-MoS2@NiS2 is higher than that of 1T-phase MoS2 or NiS2, indicating its higher electron transfer efficiency. Moreover, 1T-MoS2@NiS2 has a lower ΔGH2O (water dissociation energy barrier, 0.1 eV) and ΔGH* (−0.12 eV) than 1T/2H-MoS2 (ΔGH2O, 0.16 eV; ΔGH*, 0.97 eV) and 2H-phase MoS2 (ΔGH2O, 0.82 eV; ΔGH*, 1.49 eV), which favors the cleaving of the HO–H bond and delivers a near-neutral H* adsorption strength. Huang and coworkers constructed a MoO2/MoS2 heterogeneous nanorod encapsulated in N and S co-doped carbon (MoO2/MoS2@NSC), in which MoO2 and MoS2 were tightly coupled at the atomic scale.116 When applied to the HER in alkaline media, MoO2/MoS2@NSC exhibited an overpotential of as low as 156 mV at a current density of 10 mA cm−2 and a Tafel slope of 99 mV dec−1, which exceeded those of MoS2@NSC (232 mV, 131 mV dec−1) and S-MoO2@NSC (364 mV, 142 mV dec−1). Theoretical simulations implied that the positively charged Mo sites in MoO2/MoS2 exhibited the most negative water adsorption energy and the Mo–O* bond shortens, while the O*–H bond elongates after water adsorption, thus promoting the dissociation of H2O. In addition, the ΔGH* value of MoO2/MoS2 is significantly lower than that of MoS2, indicating the enhanced HER activity. Lang and coworkers proposed a strategy of coupling S-vacancies with single atomic Ru to enhance the HER performance of MoS2.117 Theoretical analysis showed that Ru1@D-MoS2 (i.e., atomic Ru dispersed in defective MoS2), in which Ru replaces the Mo site, has a ΔGH2O value of 0.55 eV, which is lower than those of MoS2, MoS2-Sv, Ru1(S)@MoS2, and Ru1(Mo)@MoS2, suggesting that the coupling of vacancies and Ru greatly facilitates the dissociation of water. Also, Ru1@D-MoS2 has the optimal ΔGH* of −0.18 eV, indicating its superior activity. In the experiment, the exfoliation of MoS2 and the introduction of defects were performed via ball milling and annealing. When serving as a catalyst, Ru1@D-MoS2 displayed an overpotential of 107 mV at 10 mA cm−2 and a Tafel slope value of 96 mV dec−1 in 1 M KOH solution, much better than those of D-MoS2 (∼264 mV, 85.6 mV dec−1) and commercial MoS2 (364 mV, 111.7 mV dec−1). In conclusion, the above discussion demonstrates that the defects introduced into crystalline MoS2 also contribute to an increase in HER activity in alkaline media by lowering the reaction energy barrier.

3. Defective non-two-dimensional molybdenum sulfide for the HER

Distinct from 2D crystalline 2H/1T/3R phase MoS2, non-2D molybdenum sulfides such as amorphous MoSx (α-MoSx) and clustered MoSx feature structural uncertainty, complicated internal linkage mode, lack of long-range order, and varied morphology without two-dimensional layer structure, leading to critical obstacles in unraveling the catalytic functions. Although the complex intrinsic structures sometimes cannot be fully elucidated, a few relatively well-defined cluster units, such as [Mo3S13]2−, [Mo3S12]2−, and [Mo2S12]2−, still provide a partial basis for further studies to explore the macroscopic chain/island/network morphology, underlying electronic structures, and physicochemical properties. Their superb electrocatalytic HER performance and low-cost fabrication processes have stimulated consecutive research on the synthesis of materials, as well as the identification and optimization of catalytic active sites. In the next section, we will review some representative works on amorphous MoSx and MoSx clusters for their HER application in an acidic environment. The catalyst structure, the chemical bonding transition upon activation, and the active sites play a critical role in the catalytic performance, so we closely integrate them in our discussion.

3.1. Amorphous MoSx for the HER

Amorphous MoSx, with an overpotential range of 100–180 mV at a HER current density of 10 mA cm−2 in acidic media, exhibits higher catalytic activity than highly crystalline MoS2, motivating researchers to investigate its structure–activity relationship.118–120α-MoSx, whose structure is mainly responsible for its electrochemical properties, is assigned to be a molecule-based coordination polymer consisting of cluster entities such as [Mo3S13]2−, [Mo3S12]2−, [Mo3S4]4+, [Mo2S12]2−, etc.121,122 Five sulfur ligands (Fig. 9a), including bridging S22−, apical S2−, shared S22−, terminal S22−, and unsaturated S2− (broken S22−), distinguished by their respective Raman and XPS features, stabilize isolated Mo atoms with apical S2− in the center surrounded by bridging S22−/terminal S22−, as well as shared S22−/unsaturated S2− connecting the discrete units. Together, they construct randomly disordered morphologies in the form of unfolded 1D polymeric chains or interconnected branches.122 For the detailed characterization of sulfur, the S–S stretching vibrations of terminal S22− [ν(S–S)te at 525 cm−1], bridging/shared S22− [ν(S–S)br/sh at 555 cm−1], and apical S2− [ν(Mo33S) at 450 cm−1] can be assigned separately in the Raman peaks, where ν(Mo–S)coupled is located at 382–284 cm−1 (Fig. 9b). Moreover, XPS exhibits a lower electron binding energy of 161.8 eV assigned to terminal S22−/unsaturated S2− and a higher electron binding energy of 163.1 eV assigned to bridging S22−/shared S22− or apical S2− peaks.118,122,123 In addition, the chemical state and coordination environment of Mo are affected in the amorphous state compared to the crystalline phase by, (1) the presence of a higher Mo valent state due to MoO3 or Mo[double bond, length as m-dash]O excluding common Mo4+ in Mo3IV–S, (2) the coordination number (CN) of Mo–S changes from 4.25 in the 2H phase to 5.29 and 0.60 (short) in the amorphous state, (3) the Mo–Mo distance (2.1 ± 0.7 Å) is shortened from that of the 2H phase (3.1 Å), and (4) the S:Mo stoichiometries extend to a larger range from 1.7 to 4 depending on the fabrication methods, while the oxygen atoms are concurrently preserved in the material.118,119,123,124 Based on the above structures, we will next focus on the preparation of amorphous MoSx and the identification of HER active sites, including unsaturated Mo, S22−, unsaturated S2−, and the sites synergized by Mo, S, or bonding.
image file: d3mh00462g-f9.tif
Fig. 9 (a) The α-MoSx coordination polymer with [Mo3S13]2− building block units sharing two of their three terminal disulfide bonds to form a polymeric chain. Some defective MoV[double bond, length as m-dash]O sites are present in the polymer. The apical sulfide, shared disulfide, terminal disulfide, and bridging disulfide are labeled. (b) Raman spectra of freshly electrodeposited α-MoSx and α-MoSx after equilibration in pH 7 at constant potentials of −0.45 V, −0.55 V and −0.71 V versus RHE. (c) Proposed catalytic cycle pathway for H2 evolution of α-MoSx. (a–c) Reproduced with permission.122 Copyright 2016, Springer Nature. (d) Turnover frequencies of amorphous MoSxversus the percentage of S atoms with high electron binding energy. The orange and green dots represent [Mo3S13]2− and [Mo3S4]4+ clusters, respectively. (e) H-binding free energy (ΔGH*) of the cluster, polymer and MoS2 slab at the 1-terminal, 2-terminal, 1-bridging, 2-bridging and apical sulfur sites. The adsorption conformations are shown in the panel below, where yellow, green and while balls indicate S, Mo and H atoms, respectively. (d and e) Reproduced with permission.124 Copyright 2016, American Chemical Society. (f) Raman spectra of MoSx-CE during the cathodic half sweep of cyclic voltammetry. The emerging peak of 2530 cm−1 is labeled, which is ascribed to the S–H stretching vibration of MoSx–H moieties. Reproduced with permission.123 Copyright 2016, American Chemical Society. (g) Mo K-edge Fourier transform EXAFS (k3-weighted) of 2H-phase MoS2, 1T-phase MoS2, and Am-MoS2 before (as fresh) and after (as spent) stability test. Reproduced with permission.119 Copyright 2019, John Wiley & Sons, Inc.

For the preparation of amorphous MoSx, it is important to ensure mild temperatures throughout the process. The reported methods include electrodeposition, atomic layer deposition, magnetron sputtering, wet chemical synthesis, femtosecond laser ablation fabrication, physical vapor deposition, and hydrothermal synthesis.118,125 Among them, the electrodeposition technique is used as a simple fabrication method to form amorphous films on conductive electrodes by reduction or oxidation pathways using the chemically aqueous tetrathiomolybdenum [MoS4]2− precursor in the following reactions, i.e., [MoS4]2− → MoS3 + 1/8S8 + 2e in the anodic step or [MoS4]2− + 2H2O + 2e → MoS2 + 2SH + 2OH in the cathodic step. The above deposition routes avoid the time-consuming annealing or exfoliation operations that are required in producing crystalline MoS2, endowing it with tremendous excellent scalability.122,124 Notably, an activation process is needed for the electrodeposited amorphous MoSx to show optimized activity. The activation steps eliminate the inert S ligands and expose the active species through structural transformation. For instance, an activation method can involve electrochemical cathodic scanning by linear sweep voltammetry (LSV), cyclic voltammetry (CV), or potentiostatic techniques to transform the as-deposited MoS3 to MoS2+x that resembles MoS2 but in an amorphous state and concurrently to reduce the amount of S ligands that affect the exposure and identification of the actual catalysis sites. The representative works of such activation methods are detailed in Table 5.

Table 5 Summary of the deposition technology, activation procedures, variations of Mo/S, structural transformation, identification of active sites, and HER performance of amorphous MoSx in acid
Deposition technologyRef. Activation procedure Variation of S Variation of Mo Structural transformation Active site identification HER performance after activation
Anodically electrodeposited AE-MoSx53 Reductive/ electro-oxidative/oxidative-reductive cycling Bridging S22− cleaving; terminal S22− dissolution; unsaturated S2− generation in very acidic pHs The unsaturated Mo5+ in Mo5+OxSy species are stable in mildly acidic pHs From AE-MoSx to Mo5+OxSy in mildly acid Unsaturated S2− (0 ≤ pH ≤ 2); Mo5+OxSy (3 ≤ pH ≤ 6) η = 256 mV@10 mA cm−2; over 12 h at pH 0–6
Sputtered amorphous MoSx118 Electrochemical cycling from +0.2 to −0.3 V in acid Terminal and bridging S22− almost disappear; loss of sulfur in H2S form Appearance of Mo6+ due to exposure to air From amorphous Mo3S13 or Mo3S12 entities to MoS2-like nanoislands Unsaturated Mo dangling bonds Mo-(S-□) offer sites for proton reduction η = 180 mV@10 mA cm−2; stability for over 10 h
Electrodeposition MoSx from (NH4)2MoS4 by CV121 Subjected to pre-catalytic (0.3 V) and catalytic (−0.3 V) Bridging disulfides are absent; oxidation of sulfur ligands under −0.3 V Molybdenum centers are reduced (formally Mo (III)) at −0.3 V From amorphous MoS3 to amorphous MoS2+x Interfacial MoIII(S2) units serve as putative active sites by adding H atoms to MoIV(S)(SH) species η = 250 mV@∼−4 mA cm−2 in pH 2
Electrochemical oxidation of [MoS4]2− to α-MoSx122 Constant potential of −0.71 VNHE at pH 7 Terminal S22− disappears; bridging and shared S22− decreases; net S loss Mo[double bond, length as m-dash]O motifs increase; formation of unsaturated MoIV sites From [MoS4]2− to MoS2+x MoIV–□ contributes to HER through intermediates MoIV–□, MoV–□, MoV–H, MoV[double bond, length as m-dash]O η ≈ 350 mV@−0.5 mA cm−2 in pH 1.8
Anodically deposited MoSx-AE, cathodically deposited MoSx-CE123 CV scan from 0.4 to −0.22 V Bridging and terminal S22− reduce, unsaturated S2− increases for MoSx-AE; and these for MoSx-CE maintain Appearance of MoVI /MoO3 due to oxidation Transform MoSx-AE to a structure with MoS2 composition Mo–S–H species serves as active intermediate and S is assigned to be active site η = 200 mV@∼−5 mA cm−2, Tafel slope 40 mV dec−1
Anodically deposited MoSx-AE, cathodically deposited MoSx-CE124 LSV scan cycling; post-LSV stripping to anodic potential Sulfur with high BE increases after stripping to 0.75 V N/A Irreversible reduction during LSV scan for MoSx-AE Sulfur with higher binding energy such as bridging S22− motifs TOF 0.51 s−1 from originally 0.29 s−1@η = 200 mV
Femtosecond laser ablation (NH4)2MoS4 to a-MoSx125 CV for 1000 cycles Terminal S22− disappears; unsaturated, bridging (di)sulfide reduce; oxidized sulfur increases Decrease of Mo5+ state while increment of Mo4+, Mo6+ states N/A May be bridging S22−, apical S2−; or more likely to be MoV defect by forming MoIV-□ center η = 145 mV@10 mA cm−2; stability over 7000 s
CVD deposition at 300 °C126 LSV cycles Apical S2−and/or bridging S22− nearly disappear, terminal S22−and/or unsaturated S2− maintain Appearance of higher oxidation state of Mo (Mo6+ and Mo5+) From MoS3 to amorphous MoS2 under negative potential Terminal S22− and/or unsaturated S2−, and more likely to be terminal S22− by DFT η ≈ 175 mV@10 mA cm−2, Tafel slope ∼40 mV dec−1
Deposition of MoS3-AE and MoS2-CE by potential cycling129 Linear sweep at potentials only more positive than HER Reductive removal of sulfide below one equivalent; the ratio of S22− decreases Appearance of Mo state corresponding to MoO3 for activated MoS3 To form MoS2+x from MoS3 N/A η ≈ 170 mV@20 mA cm−2 at 0.2 mg cm−2


To determine the active site of the activated α-MoSx, previous work has tracked changes in Mo and S elements by employing in situ or ex situ measurements. To date, the detailed mechanism remains elusive, although unsaturated Mo, S22− and unsaturated S2− have been considered as active centers according to individual studies. Tran and coworkers proposed unsaturated Mo to be the active center.122 In their study, Mo–□ was created by reductive activation according to the following hypothetical stepwise or concerted processes: Mo(μ-S2)Mo + 2e + xH+ → Mo([double bond, length as m-dash]S)2−x(SH)x + Mo–□, or Mo(S2) + 2e + 2H+ → Mo–□ + HS, when a constant potential of −0.71 VNHE was applied in neutral media. Disulfide ligands including bridging, shared, and terminal S22− were significantly decreased or even disappeared during the activation process, indicating that the disulfide would not be able to support a continuous HER. In addition, the unsaturated MoIV-□, which is regarded as the catalytic species, can be converted to MoV-□ after protonation and H2 evolution steps, and the hydration of this MoV-□ yields a Mo[double bond, length as m-dash]O motif which can be determined by the enhanced resonance Raman signal after activation treatment, further suggesting that the active site is indeed from unsaturated Mo (Fig. 9c). Moreover, the much lower H adsorption free energy ΔGH* (0.108 eV) to form MoV-hydride species from MoV-□ also favor the mechanistic route via the Mo–H intermediate prior to H2 evolution. Xi and coworkers also concluded that the unsaturated Mo center is the active site, where a similar S22− reduction route of MoS2 + 2e + 2H+ →Mo–(S–□) + HS is considered to generate unsaturated Mo dangling bonds.118 This hypothesis is supported by the complete disappearance of the terminal and bridging S22− as shown by online Raman spectroscopy and the ratio of Mo6+ to Mo4+ state increasing to 1[thin space (1/6-em)]:[thin space (1/6-em)]3.14 due to the susceptibility of unsaturated Mo atoms with dangling bonds to oxidation in air. More specifically, S22− has been partially transformed into S2− existing in the emerging MoS2 nanoislands converted from α-MoSx, while the other sulfur losses have been observed to be released as H2S gas, especially in the early stage of the activation process. The exposed Mo–(S–□) active species, together with its Mo–(S–H*) intermediate adsorption state, confer excellent HER activity to the activated catalyst with an overpotential of 180 mV at a current density of 10 mA cm−2 and a stability test over 10 h in a 0.5 M sulfuric acid electrolyte, demonstrating the significant role of the activation process in this amorphous catalyst. Li and coworkers also concluded that MoV defective species are the active sites and the MoV species can be controllably generated by temporally shaped femtosecond laser ablation of (NH4)2MoS4 aqueous solution through tuning the laser energy and pulse delay to control reduction/oxidation.125 This work revealed a close relationship between the HER activity in acidic media and the MoV species content, showing an incremental HER activity with optimized onset potentials (127–105 mV), overpotentials (173–145 mV) at a current density of 10 mA cm−2 and Tafel slope values (47–40 mV dec−1), with increasing MoV state content (18.2–28.3%). However, after the CV activation process, the content of the MoV state decreased, while the content of MoIV increased; therefore, this work assumed a similar transition from MoIV to unsaturated catalytic MoIV–□ centers. Unfortunately, the corresponding studies still lack the direct evidence to capture such MoIV–□ catalytic species responsible for hydrogen evolution.

The sulfur ligands are also designated as the active site. Among the five types of apical S2−, bridging S22−, terminal S22−, shared S22−, and unsaturated S2−, terminal S22−, bridging S22−, and unsaturated S2− are successively considered as the actual catalytic sites. Li and coworkers proposed that the terminal S22− may undertake the catalytic function in the amorphous MoSx deposited by CVD at low temperatures.126 In brief, this amorphous material grown at 300 °C contained abundant apical S2− and/or bridging S22−, as well as terminal S22− and/or unsaturated S2−, but after LSV cycling, the apical and/or bridging S22− nearly disappeared as shown by XPS. Thus, the high HER activity that was stable during LSV cycling should come mainly from terminal S22− and/or unsaturated S2− and the authors particularly emphasized the vital role of terminal S22− in the catalyst, as suggested by the DFT results. In contrast, another investigation conducted by Ting and coworkers concluded that bridging S22− may be the major catalytic center towards proton reduction.124 The bridging S22− was generated by an activation method designed to perform CV oxidative stripping to an anodic potential of 0.75 VRHE, leading to an increase in the proportion of sulfur with higher electron binding energy due to oxidation of sulfur with lower binding energy, based on post-LSV MoSx-AE films (MoSx electrodeposited on the anode). To measure the catalytic activity, TOF measurement instead of LSV curves was carried out to exclude the possible influence from different catalyst loadings. At an overpotential of 200 mV, a linear positive correlation was established between the percentage of S atoms with higher electron binding energy (e.g., bridging S22− and apical S2−) and TOF (see Fig. 9d for details), while apical S2− was excluded from the main catalytic reaction sites due to its weak H adsorption (ΔGH* > 1 eV, Fig. 9e), suggesting that the bridging S22− site in the catalyst facilitated the evolution of H2.

Additionally, according to Deng and coworkers, unsaturated S2− may play a dominant role in the evolution of H2 when more negative potentials are applied to the catalyst.123Operando Raman spectroscopy recorded detailed changes during the CV cathodic sweep at a slow scan rate of 0.5 mV s−1, showing the attenuation of apical S2−, terminal S22−, and bridging S22− at −0.07 V. This suggests that the S–S bonds of the terminal and bridging S22− are cleaved before being converted to possible unsaturated S2− species, consistent with the XPS results showing an increase in the proportion of S atoms with low electron binding energy (terminal S22− and unsaturated S2−) after the HER. More importantly, a peak at 2530 cm−1 belonging to the S–H stretching vibration was captured at potentials below −0.07 V (Fig. 9f) and its signal increased with more negative potentials and faded away when the potential sweep was reversed to positive values, indicating the appearance of MoSx–H species under acidic HER conditions. The single S atom bound to H is thought to be responsible for this HER activity although the type of sulfur cannot be assigned accurately. Consistently, the absence of Mo–H signals further pinpoints the sulfur atoms as the active sites.

Accordingly, since the different studies mentioned above point to different active sites and give corresponding experimental and theoretical evidence, we try to believe that the S active sites may depend on the preparation and the activation methods of the material, since the internal structures of amorphous MoSx prepared by different processes are clearly different. We also see that the accurate identification of active sites remains a challenge due to the lack of techniques to individually identify each sulfur ligand or Mo intermediate during the catalytic reaction.

Beyond the above discussion on individual S or Mo elements serving as active sites, Lassalle-Kaiser and coworkers proposed a synergistic effect arising from Mo and S cooperation.121 In their study, MoIV was reduced to the MoIII state according to the Mo L-edge X-ray absorption near energy structure (XANES) results, while S was oxidized to the terminal S22− ligand according to the S K-edge XANES results, which occurred at the solid–liquid interface of MoS2+x at −0.3 V (vs. RHE). The MoIII(S2) units, considered to be catalytically active intermediates, underwent protonation and reduction steps to form MoIV(S)(SH) species prior to releasing H2 and renewing the intermediate, exhibiting their cooperative involvement in catalysis. Escalera-López and colleagues combined XPS, Raman and electrochemical tests to reveal the pH-dependence of the active sites in anodically electrodeposited amorphous molybdenum sulfide (AE-MoSx).53 In a very acidic environment (0 ≤ pH ≤ 2), the predominant HER active sites were unsaturated S2− after bridging S22−cleaving and terminal S22− dissolution at the cathodic potential. In contrast, under mildly acidic conditions (3 ≤ pH ≤ 6), the Mo5+OxSy species (i.e., oxygen-containing and S-deficient Mo-sites) with unsaturated Mo5+ were thermodynamically more stable on the AE-MoSx surface and thus successfully catalyzed the HER. Thus, relying on the two active species, the decent stability was observed over a wide range of acidic pHs (0–6) in a 12 h (at 10 mA cm−2) stability test. Moreover, besides the element itself, the bond structure in amorphous phase MoSx is also considered to play a key role in determining the activity. According to a recent study by Wu and coworkers, the inherent short Mo–Mo bonds in the amorphous phase compared to the 2H or 1T phases (Fig. 9g), were elucidated to enhance HER activity and maintain its stability, enlightening us to recognize its active center from a systematic perspective combining materials and structures.119 Especially for doped amorphous MoS2+x, such as N-doped α-MoSx that displayed a smaller overpotential than α-MoSx as reported by Ding and coworkers, it poses a new challenge to distinguish the active center since its complicated components require a systematic study of more factors, including the types of elements, the occupied sites of exotic atoms, the ligand types, and the bond structures, to draw conclusions.127 Certainly, in addition to the above work on active site identification, it is equally important to focus on the performance breakthroughs. Zhang and colleagues prepared a W-doped α-MoWSx/N-RGO catalyst, which achieved a low overpotential of 348 mV and 24 h stability at an industrial-level current density of 1000 mA cm−2 in an acidic electrolyte and the overpotential of this catalyst was lower than that of Pt/C at high current densities, indicating that the excellent performance of α-MoSx can be obtained through rational structural design.128

Undoubtedly, the cost-effective, efficient, and facile synthesis process by electrodeposition or wet chemical routes under ambient conditions makes such amorphous materials attractive for large-scale H2 evolution in the PEM electrolyzer. Further investigations are worthwhile to, (1) identify the active sites with advanced techniques, (2) balance amorphous and crystalline states for better activity and stability, and (3) scale up controllable manufacturing methods.

3.2. MoSx clusters for the HER

The molecular monomers of amorphous MoSx (e.g., Mo3S132− cluster) can independently bear hydrogen production, benefiting from the structural motifs resembling edge active sites in MoS2. MoSx clusters present a rich variety of cluster types and structures, including the basic Mo3S132−, Mo3S122−, MoS42−, Mo2S122−, Mo3S44+, etc., which contain only Mo and S elements. Moreover, the incorporation of exotic elements or molecular ligands results in more complex structures and modifiable HER properties. In the next section, we will first discuss several basic cluster structures and their HER performance, including Mo3S132−, Mo3S122−, and Mo2S122− clusters, followed by the modification methods of MoSx clusters, such as oxygen incorporation, ligand linkage, and heteroatom doping; notably, in each cluster, the preparation method, structure, and properties are in focus.

Among these clusters, the Mo3S132− cluster is considered to be the most representative entity to show prominence in the catalytic HER. Kibsgaard and colleagues elucidated that Mo3S132− consists of three terminal S22− with lower binding energy, three bridging S22− and one central apical S2− with higher binding energy (Fig. 10a), with a ratio of 7[thin space (1/6-em)]:[thin space (1/6-em)]6 between the doublet with higher binding energy and the other doublet.55 The Mo3S132− ion cluster could be stored in the synthetic (NH4)2Mo3S13·nH2O, a highly crystalline dark red bulk that can be easily deposited on the substrate by the wet-drop casting method, and the individual cluster can be imaged by atom resolved STM imaging (Fig. 10a), showing an ordered atomic-level structure. This catalyst with loadings of 10–100 μg cm−2 exhibited onset overpotentials of 0.1–0.12 V, low overpotentials between 0.18 V and 0.22 V at a current density of 10 mA cm−2, superb Tafel slope values of 38–40 mV dec−1 under acidic conditions, and comparable or better TOF values than amorphous MoSx (Fig. 10b).


image file: d3mh00462g-f10.tif
Fig. 10 (a) Top and side views of a single [Mo3S13]2− cluster model (left panel) and atomically resolved STM image of a single [Mo3S13]2− cluster on a highly orientated pyrolytic graphite surface, showing the ordered atomic-scale structure (right panel). (b) TOF of [Mo3S13]2− clusters with different loadings and substrates, as well as TOF of several other HER catalysts based on MoSx. (a and b) Reproduced with permission.55 Copyright 2014, Springer Nature. (c) Hydrogen evolution activity (TOF per Mo atom at η = 200 mV) versus DFT-calculated ΔGH* for [Mo3S13]2−|C, [Mo3S13]2−|Au, [Mo3S13]2−|Ag and [Mo3S13]2−|Cu electrodes. The schematic of [Mo3S13]2− and metal bulk is shown in the inset. Reproduced with permission.130 Copyright 2017, American Chemical Society. (d) Left panel: Schematic of the full cell setup with the cathode, Nafion PEM, and anode. The catalyst is sprayed onto a carbon cloth porous transport layer to form porous transport electrodes. Right panel: Polarization data of the pristine membrane electrode assembly (MEA) and the same MEA after 100-h testing on a Mo3S13 or Pt catalyst. Reproduced with permission.131 Copyright 2020, John Wiley & Sons, Inc.

It is noteworthy that in the above study, the two substrates, HOPG and GP, are loaded with sub-monolayer and bulk catalysts, respectively, resulting in a disparate Tafel slope and TOF values, as well as the underlying reaction mechanism pathways.55 In another similar study by Hellstern and coworkers, changing the substrate produced considerable differences in activity even at similarly low coverages of the catalyst on the substrate, where the authors have ensured the proximity of the catalyst to the substrate.130 In their study, a comparative study of copper, glassy carbon, gold, and silver substrates was performed to demonstrate the correlation between the calculated ΔGH* and TOF values (Fig. 10c), where a gold substrate loaded with Mo3S132− achieved an optimal TOF of 0.47 H2 s−1 per Mo atom (in an acidic electrolyte) and also an optimized thermal adsorption/desorption energy. Thus, the choice of substrates has a great influence on the HER activity, which comes from the possible chemical rearrangement step, electrical conductivity, support affinity, active surface atom from substrate, etc., thus providing a perspective to modulate catalytic activity by substrate engineering, especially in the case of low coverage of clusters.

The Mo3S132− cluster described above was synthesized via a wet chemical route, followed by deposition on the scaffold by simple drop-casting or spray coating methods, allowing them to be easily scaled up and applied in PEM electrolyzer devices. Holzapfel and coworkers recently conducted a feasibility study on the application of this material in PEM devices, where Mo3S132− clusters and a N-doped carbon nanotube (NCNT) hybrid were used as the cathode and IrO2 as the anode, which were separated by a Nafion membrane.131 The electrolyzer achieved a high current density of 4 A cm−2 at a cell voltage of 2.36 V under a catalyst load of 3 mg cm−2 (Fig. 10d), which surpassed the current density of the corresponding non-precious metal catalyst in the PEM cell, and attained a slight degradation of 83 μV h−1 in a 100 h stability test at 1 A cm−2. In addition, component dissolution experiments involving cluster structure stability were performed by inductively coupled plasma mass spectrometry (ICP-MS) at downstream of a three-electrode electrochemical setup, and it was found that the Mo element readily dissolved upon contact with the electrolyte, but then the dissolution slowed down rapidly. This slight dissolution could be ascribed to the formation of Mo6+ oxides during air exposure although the balance between Mo6+ oxides and Mo3S132− is uncertain, however, the Mo3S132− clusters originally underneath the oxide layer were capable of maintaining catalytic activity in the PEM electrolyzer after the depletion of the oxide layer.

Similar to Mo3S132− in structure and composition, Mo3S122−, Mo2S122−, and [Mo3S4]4+ also possess high HER activity, which enriches the structural variety of the clusters. The counterions of these cluster ions are usually NH4+ or SO42−, whose effects on catalysis remains to be investigated.132,133 Simple chemical preparation methods are also applicable to them; for instance, Mo2S122− can be obtained in as-synthesized crystalline (NH4)2[Mo2(S2)6] 2H2O with orthorhombic space group Pnna, where each Mo center is coordinated to two bridging S2 and two terminal S2;134 and cubane-type [Mo3S4]4+ could be produced by a simple (NH4)2MoS4 reduction reaction.133 To obtain Mo3S122−, Lee and colleagues developed a transformation procedure that revealed that the Mo3S132−, Mo3S122−, and low-crystallinity MoS2 could be obtained through the reaction of MoS3 and ammonium hydroxide solution by eliminating the apical S and the bridging S below and above 280 °C, respectively.135 In the case of Mo3S122−, the current density at each potential is slightly inferior to that of Mo3S132− and the bridging S is considered as the active site of Mo3S122−, as suggested by the decrease in current density after elimination of the bridging S. In another work, Mo2S122− with a loading of 150 nmol cm−2 exhibited a better polarization profile than Mo3S132− or other cluster-type catalysts (Fig. 11a) with an overpotential of 161 mV at 10 mA cm−2 and a Tafel slope of 40 mV dec−1 in acidic solution.134,135 The complementary theoretical study revealed that Mo2S122− preferentially follows the Volmer–Heyrovsky reaction pathway because the activation energy barrier along the Volmer–Tafel reaction is larger (Fig. 11b), supporting the experimental Tafel slope (40 mV dec−1) and thus the limiting step comes from the electrochemical desorption process.134


image file: d3mh00462g-f11.tif
Fig. 11 (a) Comparison of the overpotential of various MoSx-based catalysts at a HER current density of 10 mA cm−2. (b) Catalytic reaction pathways and energies of the HER on Mo2S12 with two different reaction mechanisms (Volmer–Heyrovsky and Volmer–Tafel reaction). (a and b) Reproduced with permission.134 Copyright 2015, John Wiley & Sons, Inc. (c) Thermodynamic route of compound [Mo2O2(μ-S)2(S2)2]2− with two variables: electron addition (vertical) and proton addition (horizontal). Series of isomers and the energies of each step are shown. Reproduced with permission.136 Copyright 2019, Springer Nature. (d) Synthesis and linking of Mo3S7 clusters by ligand substitution. The blue, orange, yellow, and purple balls represent Mo, Br, S, and groups, respectively. Reproduced with permission.137 Copyright 2018, American Chemical Society.

The tuning of the cluster structure confers flexibility and variability to their structures, thus providing many opportunities for breakthroughs in catalytic performance. At least three strategies, namely oxygen incorporation, ligand linkage, and heteroatom doping are effective in optimizing the cluster-based HER activity. McAllister and coworkers fabricated two oxygen-incorporated clusters, [Mo2O2(μ-S)2(S2)(S4)]2− and [Mo2O2(μ-S)2(S2)(S2)]2−, by a wet chemical method.136 In these clusters, in addition to the Mo[double bond, length as m-dash]O bond, the Mo centers are also mutually connected by two sulfur bridges and each Mo center is coordinated to a terminal disulfide (S2) or tetrasulfide (S4) group. The catalyst achieved a very low overpotential of about 114 mV at a current density of 10 mA cm−2, a marginal change of the observed overpotential after 1000 cycles in 1 M H2SO4 media, as well as Tafel slope values of 52–55 mV dec−1 signifying an efficient Heyrovsky pathway. The simulation results revealed that the Heyrovsky step proceeds along the reaction of H2[Mo2O2(μ-S)2(S2)2]2− + H5O2+ → H[Mo2O2(μ-S)2(S2)2] + 2H2O + H2 with a reaction Gibbs free energy (ΔG°) of −7.3 kcal mol−1, where the H[Mo2O2(μ-S)2(S2)2] anion is considered as the catalytic cycle intermediate and H2[Mo2O2(μ-S)2(S2)2]2− is formed via electron addition and proton addition pathways (Fig. 11c).

Except for the coordination of pure oxychalcogenide to the Mo center, organic molecules can also act as the ligands inside cluster-like complexes such as MoO(S2)2(2,2′-bipyridine) and [MoO(S2)2picolinate],138,139 or act as linkers among individual clusters to build larger entities. Ji and colleagues developed a linkage method to organize single molecular clusters into ordered dimers, cages, and chains through sulfido and thiolate linkers.137 This process commenced with the [Mo3S13]2− cluster that underwent bromide ion replacement on terminal S2 sites followed by further ligand substitution into thiolate [Mo3S7Br3(SR)3]2− (Fig. 11d), where 1,3-benzenedithiol yields the dimer (NEt4)4[Mo3S7Br3(m-BDT)1.5]2 (MOS-1) and 1,4-benzenedithiol produces the dimer (NEt4)4[Mo3S7Br4(p-BDT)]2 (MOS-2) or (NEt4)2[Mo3S7Br3(p-BDT)1.5] (MOS-3) depending on the orientation of clusters. MOS-3 delivers a low overpotential of 89 mV in driving a current density of 10 mA cm−2 (see Table 6 for the comparison of the HER performance of MoSx clusters), with a Tafel slope of 57 mV dec−1, which are better than those of [Mo3S13]2− in acidic media, demonstrating the key role of ligand modification. Additionally, Escalera-López and coworkers enhanced the HER properties via a heteroatom doping strategy, such as nickel-doped MoSx nanoclusters.54 Nickel element can be introduced into MoSx by dual-target magnetron sputtering deposition to obtain (Ni–MoSx)1000 clusters, which outperform undoped (MoSx)300 or Ni2200 clusters in the proton reduction reaction in acid media, also extending the modification methods and diversity of the clusters.

Table 6 Summary of HER performance of MoSx-based cluster catalysts in acid
ClusterRef. Preparation process Electrolyte Overpotential (vs. RHE) Tafel slope (mV dec−1) Stability Loading TOF/exchange current density (j0)
(Ni–MoS2)1000 nanocluster54 Dual-target magnetron sputtering deposition 2 mM HClO4/0.1 M NaClO4 640 mV@0.31 mA cm−2 122 N/A 4.25 μg cm−2 TOFMo-edge site 60.3 s−1@η = 0.74 V
J0 2.1 × 10−9 mA cm−2
[Mo3S13]2−|GP55 Wet-chemistry methods 0.5 M H2SO4 180 mV@10 mA cm−2 40 Slight change after 1000 cycles 50 μg cm−2 TOFper Mo atom 0.5 s−1@η = 0.2 V
[Mo3S13]2−|Au130 Wet-chemistry synthesis 0.1 M HClO4 ∼266 mV@5 mA cm−2 58 N/A 3.59 × 1015 Mo atoms cm−2 TOFper Mo atom 0.47 s−1@−0.2 V
Mo3S13-NCNT131 Self-assembly process 0.5 M H2SO4 188 mV@10 mA cm−2 40 100 h of full cell 3 mg cm−2 at cathode N/A
[Mo3S4]4+ sub-monolayer133 Chemical reduction 0.5 M H2SO4 ∼400 mV@0.1 mA cm−2 120 Decrease after 10 sweeps for multilayer 1.6(± 0.2) × 10−11 mol cm−2 TOFper molecule 0.07 s−1
J0 2.2 × 10−7 A cm−2
[Mo2S12]2− cluster134 A modified Müller method 0.5 M H2SO4 161 mV@10 mA cm−2 39 Marginal loss after 1000 cycles 10/50/150 nmol cm−2 TOFper Mo atom 3.27 ± 0.15 s−1@η = 0.2 V
Mo3S13/O-CNT135 Reaction of α-MoS3 and NH4OH 0.5 M H2SO4 137 mV@10 mA cm−2 40 1000 cycles 54 μg cm−2 TOFper Mo atom 2.5 s−1@−0.25 V
[Mo2O2(μ-S)2(S2)2]2− cluster136 Wet-chemical methods 1 M H2SO4 114 ± 3 mV@10 mA cm−2 52 <18 mV shift at 10 mA cm−2 after 1000 cycles 2.85 μmol cm−2 TOFper active site 0.12 s−1@η = 0.2 V
(NEt4)2[Mo3S7Br3(p-BDT)1.5]137 Ligand substitution 0.5 M H2SO4 89 mV@10 mA cm−2 57 ∼37% decay after 5000 s@η = 89 mV 2.7 μmol cm−2 TOFper Mo atom 0.48 s−1@η = 0.2 V
J 0 2.5 mA cm−2


It bears some similarities among various amorphous MoSx in terms of facile growth conditions, scalable wet chemistry, and high catalytic efficiency of the clusters; moreover, the flexibility and diversity of composition and linkage structures endow it with more opportunities to make superior HER catalysts with high performance, low cost, and great potential for practical applications. However, the identification of active sites in various clusters remains elusive and the stability of cluster structures remains suboptimal. Therefore, we believe that more attention should be paid to the following aspects, including the identification of active sites by advanced characterization techniques, the improvement of the catalyst durability by screening internally stable ligands, the exploration of the conversion/degradation mechanism of clusters during the hydrogen evolution process, and the enhancement of catalyst stability under continuous bubble impingement in acid by optimizing the growing integration of the clusters with the substrate. In conclusion, MoSx clusters, like defective 2H/1T/3R-phase MoS2 and amorphous MoSx, also exhibit superior performance in catalyzing the HER (see Fig. 12 for the comparison of overpotentials of these defective MoSx) and are potential alternatives to the noble metal catalysts in PEM electrolyzers.


image file: d3mh00462g-f12.tif
Fig. 12 Comparison of the HER performance of a few representative defective MoSx catalysts using stability versus overpotential at 10 mA cm−2 in acidic media.

4. Other catalytic applications of defective molybdenum sulfide

Molybdenum sulfide has demonstrated potential for catalytic applications in other areas beyond the HER, such as the NRR, CO2RR, ORR, OER, nitric oxide (NO) reduction reaction, nitrate ion reduction reaction, organic molecular upgrading, etc. and as a catalytic electrode material for energy storage devices such as the lithium-oxygen batteries, lithium-sulfur batteries, sodium–sulfur batteries, zinc–air batteries, and so on. Recently, He and coworkers proposed the “self-gating” phenomenon, in which the electrocatalytic reactions of a semiconductor catalyst can modulate the surface conductivity itself, where n-type semiconductors support reduction reactions at negative potentials, such as the HER and CO2RR; p-type semiconductors tend to support reactions at positive potentials, such as the ORR and OER; and bipolar semiconductors typically support both reactions (Fig. 13a).140 Thus n-type MoS2 is suitable for the HER and CO2RR, while p-type MoS2 such as phosphorus-doped MoS2 tends to support both ORR and OER reactions, and Ta or Nb modulated bipolar MoS2 materials are suitable for both the HER and ORR, thus guiding catalytic applications of MoSx by tuning its semiconducting properties. Certainly, for a catalytic reaction, it is insufficient to consider the conductance of catalysts alone; more aspects such as the geometric structure, electronic state, adsorption energy, reaction energy barrier, and orbital hybridization also profoundly affect the performance of catalysts and need to be considered comprehensively. In the following, we will screen and describe several representative catalytic applications of defective MoS2 in these electrocatalytic fields.
image file: d3mh00462g-f13.tif
Fig. 13 (a) Schematic of the correlation between the type of semiconductor and the preferred electrocatalytic activity. The n-type and bipolar semiconductors can be turned on by negative electrochemical potentials, making them suitable for cathodic reactions, while the p-type and bipolar semiconductors can be turned on by positive electrochemical potentials, making them suitable for anodic reactions. The VB (valence band) and CB (conduction band) are marked. Reproduced with permission.140 Copyright 2019, Springer Nature. (b) Average NH3 yields (left y-axis) and FEs (right y-axis) of MoS2/CC at negative potentials. Reproduced with permission.24 Copyright 2018, John Wiley & Sons, Inc. (c) NRR free-energy profiles for MoS2 edges with (blue) and without (red) Li–S interactions. The inset is the deformation charge density of *N2 at the MoS2 edge with Li–S interactions, where yellow and blue represent the accumulation and loss of charge, respectively. Reproduced with permission.34 Copyright 2019, John Wiley & Sons, Inc. (d) Schematic of selective electrocatalytic NH3 synthesis of Li intercalated 1T-phase MoS2. Enhanced NRR and suppressed HER are achieved due to additional Li–S bonds during Li+ intercalation. The pseudo six-membered ring containing the interaction N–Li–S–Mo–S–Mo is displayed. Reproduced with permission.144 Copyright 2021, Royal Society of Chemistry.

4.1. Nitrogen reduction reaction

The Haber–Bosch process is developed industrially for the artificial conversion of dinitrogen to ammonia, the main source of nitrogen for the production of fertilizers, whereas the process consumes 1–2% of the world's energy supply and contributes about 1.5% of global greenhouse gas emissions.33 The electro-synthesis of ammonia offers a promising solution to address these environmental and energy issues by replacing harsh conditions (650–750 K and 50–200 bar)141 with ambient conditions and a renewable supply of clean energy. Breakthroughs toward electrocatalytic N2 reduction (N2 + 6H+ + 6e → 2 NH3) have been made with Mo-based electrocatalysts, particularly MoS2. Zhang and coworkers proposed that the edges of 2D MoS2 are active for the NRR that is similar to the HER, however, the inert basal plane does not favor the reaction according to DFT calculations, which further suggests that the positively charged (+ 0.963[e]) Mo edges overcome a lower barrier to polarize and activate the N2 molecule by forming N–Mo bonds to weaken the solid N[triple bond, length as m-dash]N triple bond.24 Intrinsically, two-dimensional MoS2 with high crystallinity achieves only 1.17% FE and NH3 yield of 8.08 × 10−11 mol s−1 cm−1 in 0.1 M Na2SO4 (Fig. 13b), while the defect engineering on MoS2 greatly facilitates the electrosynthesis of ammonia. To date, strategies such as bulk structural defects, crystalline phase modulation, engineering of S site, exotic atomic incorporation, and hybridization method, which will be discussed sequentially in the next section, have been proven to be effective for activating the NRR process, suppressing the HER reaction, and thus enhancing the product selectivity.

For direct bulk structural defects, Li and coworkers developed a defective MoS2 nanoflower with dislocations and distortions on the crystal, leading to disordered atomic arrangement and lattice expansion on the defect sites.142 These defects and disorders with abundant unsaturated atoms boosted the catalytic activity of the NRR with NH3 yield and FE of 29.28 μg h−1 mg−1 and 8.34%, respectively, much better than those of MoS2 with fewer defects (13.41 μg h−1 mg−1 and 2.18%) under the same conditions. Such enhancements benefit from the lower energy barrier of the potential determining step (i.e., *NH2 → *NH3) along bare Mo atoms on the rim of defects, and also from the optimized d-band center moving closer to the Fermi level that indicates a stronger interaction between the catalyst surface and N2 molecule. Chen and colleagues designed a porous atomic layered 2H-phase MoS2 in which porosity induced the generation of adjacent Mo atom pairs.143 The Mo atom pairs effectively reduced the thermodynamic overpotential of the NRR and assisted in obtaining a high FE of 44.36% and a sizeable NH3 yield of 3405 μg h−1 mgcat−1. Wu and coworkers synthesized a sub-monolayer MoS2−x structure with in-plane S-vacancies, which tune its catalytic performance through adjusting its affinity and stability with nitrogen intermediate in a dynamic binding paradigm, thus circumventing the scaling relationship and obtaining excellent electrocatalytic NRR performance at ultra-low overpotential.145

The crystalline phase affects the NRR reaction path and HER reaction free energy; so, the catalytic performance can be tuned by phase modulation. Liu and coworkers prepared 1T/2H mixed-phase MoS2 with tunable 1T content using a self-sacrificing template method.146 The current density, NH3 yield rate, and FENRR gradually increased with increasing 1T-phase content (from ∼10% to ∼75%), while the FEHER decreased. DFT analysis showed that the Mo-edge of 1T-phase MoS2 had lower N2 adsorption energy than that of 2H-phase MoS2. Moreover, the low ΔGH* value (−1.9 eV) at low hydrogen coverage (<25%) indicates that the adsorbed H readily binds to S atom in the basal plane of 1T-phase MoS2, while the barrier for H desorption/combination cannot be overcome at the electrode potential of image file: d3mh00462g-t2.tif (−0.36 V, critical potential for overcoming the barrier of the potential-determining step). The *H remaining at the edge of the basal plane can be directly transferred to nearby bound N2 or nitrogen intermediates, which greatly accelerates the NRR process. Therefore, although the 1T-phase MoS2 can accelerate the NRR and HER, it will reduce the competition between the HER and NRR due to the separation of active sites, thus synchronizing the selectivity and activity for the NRR. Lin and colleagues compared three crystalline phases of 2H, 1T′ (zigzag chainlike clustering of Mo atoms) and 1T′′′ (trimerization of Mo atoms).147 The Mo–S distortion and Mo–Mo metal in 1T′′′-phase MoS2 raise the local electron density surrounding the Mo–Mo chains and lead to the uplift of the d band for Mo, which facilitates the acceptance of lone-pair electrons from N2 and the electron transfer from the d band to N[triple bond, length as m-dash]N π antibonding orbital, thus enhancing the adsorption to N2. In the 1T′′′ phase, the 1T′′′-3 site (the region surrounded by three Mo–Mo bonds) reinforces the electron loss of N[triple bond, length as m-dash]N triple bonds and facilitates the N[triple bond, length as m-dash]N breakage prior to the formation of a strong Mo–N σ bond. As a result, the 1T′′′-phase MoS2 catalyst delivered an NH3 yield rate of 9.09 μg h−1 mg−1, which is 2 and 9 times higher than those of 1T′ and 2H phases, respectively, as well as a FE of 13.6%.

The active sites of the NRR originate from the bare Mo atom, while the HER process arises from the S-edge primarily and also from the Mo-edge sites.34 Competition with the HER is inevitable because the HER bears a lower energy barrier and a faster reaction rate, thereby a higher NRR FE can be obtained by suppressing the HER on S-edge sites. Engineering of S sites such as interaction with Li, or substitution with F, has demonstrated strong HER suppression.34,144,148 Liu and coworkers disclosed that the in-operando created strong Li–S interaction on S-rich MoS2 nanosheets during the NNR in a lithium-containing electrolyte led to superior NRR catalytic activity and significant HER suppression, delivering an NH3 yield of 43.4 μg h−1 mg−1MoS2 and a FE of 9.81%, which are more than 8 and 18 times higher than those without the Li–S interaction (5.35 μg h−1 mg−1MoS2; 0.53%).34 Theoretical analysis revealed that the edge sites of MoS2 are electrochemically active for the NRR compared to the inert basal plane, while the Li–S interaction is thermodynamically more inclined to occur at the S-edge sites. The strong Li–S interaction significantly suppresses the HER by retarding the HER process on S-edge sites due to its uphill ΔGH* from 0.03 to 0.47 eV and by impeding the Heyrovsky/Tafel process due to downhill ΔGH* from 0.09 to −0.72 eV on Mo-edge sites. This ensures firm H* adsorption on Mo and also leads to the positively charged Mo edge sites with dramatically increased N2 adsorption energy (ΔGN2 changes from −0.32 to −0.70 eV) and a reduced NRR activation energy barrier for the formation of *N2H and other intermediates (Fig. 13c).

Moreover, Patil and coworkers explored the intercalation of Li on 1T-phase MoS2 (Li-1T-MoS2), which exhibits stronger N2 adsorption compared to Li intercalated 2H-phase MoS2 (Li-2H-MoS2), as the former possesses shorter d(Mo–N) and d(N–Li) distances according to structural calculations, which suggests a pseudo-six-membered ring containing the interaction from N–Li–S–Mo–S–Mo.144 Li-1T-MoS2 enhances the N2 adsorption on the Mo active sites with upgraded N2 adsorption energy to −0.72 eV from −0.28 eV of 1T-phase MoS2 and suppresses the HER on S atoms since the theoretical ΔGH* increases from 0.03 eV to 0.49 eV after Li intercalation (see Fig. 13d for the structure diagram). Experimentally, a high FE of 27.66% was realized from Li intercalated low-crystalline 1T-phase MoS2 formed during the NRR in the LiClO4 electrolyte. In addition to the Li interaction/intercalation engineering methods mentioned above, the F substitution on the S site can also achieve the suppression of the HER and increase NRR FE, as reported by Liang and coworkers.148 The smaller size and higher electronegativity of F, and the compression of the MoS2 interlayer space caused by the introduction of F, play a role in restraining the HER, as evidenced by the decrease in the current density of hydrogen evolution compared to pristine MoS2. Another aspect is that the Mo sites with empty d orbitals on the MoS2 edge acquire stronger N2 adsorption since the free energy change on the F-MoS2 edge (−0.85 eV) is lower than that of MoS2 (−0.38 eV) and bears the lower uphill free energy (0.36 eV vs. 0.64 eV) in the first hydrogenation step, which is considered as the NRR rate-determining step (*N2 → *NNH). As a result, the FE of the NRR on F-MoS2 increased to 20.6% with a NH3 yield of 35.7 μg h−1 mg−1.

The incorporation of exotic atoms into the MoS2 lattice creates many defects such as hetero-interfaces or S-vacancies, and the synergistic effect leads to an increase in catalytic performance. Zhao and coworkers plotted the limiting potentials of the HER and the NRR versus the N2H adsorption energy (potential-determining step) and predicted that the onset potentials of the NRR increase in the following order, Mo < Ru < V = Rh < Fe ≈ Co < Cr < Ti < Mn < Ni < Sc, based on a computational model in which the transition metal atom is embedded in a MoS2 monolayer with S-vacancies and occupies the S-vacancy site.149 A similar conclusion was drawn by Yang and colleagues. In their high-throughput screening model, a single transition metal atom was anchored on the top of the Mo site and stabilized by neighboring S atoms of MoS2; and the optimal NRR overpotential, as well as selectivity, was realized for the Mo single atom compared to loading of other metal atoms.150 Dang and coworkers proposed that a transition metal atom supported on graphene and simultaneously adsorbed on the MoS2 basal plane with S-vacancies acts as a single-atom promoter (SAP) by inducing strong spin polarization on the exposed Mo atoms, which greatly facilitates N2 adsorption and polarization.151 In their designed system, Sc or Y SAPs possess favorable NRR catalytic activity and selectivity. Many experimental studies have also been conducted to prove that introduced external atoms increase the NRR catalytic activity of the MoS2 catalyst.

Among the exotic elements, the integration of Fe atoms and MoS2 stands out in catalyzing the NRR. Chen and coworkers revealed that Fe- and O-decorated edge-MoS2 (Fe1-4O@Mo-edge-MoS2) is responsible for the high performance of the NRR due to its low Gibbs free energy and high selectivity by theoretical simulations.152 Xie and coworkers theoretically demonstrated that the MoS2-supported single Fe site induces a local electric field effect that interacts with the dipole moment of NH2–NH2 species or MoS2-supported double/triple Fe sites induce synergistic effects, both of which reduce N–N bond cleavage barriers.153 Li and coworkers experimentally reported a protrusion-like Fe single atom immobilized on monolayer MoS2 (SACs–MoS2–Fe) on the top of Mo, by coordinating Fe atoms with three nearest S.154 The Fe atom with high-curvature protrusions induces an electric field that triggers interfacial polarization, and the enhancement of electric field shown by Kelvin probe AFM follows the same trend of improvement in NRR activity (Fig. 14a and b). In detail, the increased electric field shortens the Fe–N bond and reduces the interfacial barrier, allowing electron transfer to N2 and thus promoting N2 polarization. In particular, based on the interfacial polarization field, more electrons are transferred from the Fe d-orbital to the N2 π2p* antibonding orbital. As a result, the SACs-MoS2–Fe with 2% Fe single atoms delivered a high ammonia production rate of 36.1 ± 3.6 mmol g−1 h−1 with a FE of 31.6% ± 2% at −0.2 VRHE, as well as excellent durability for 50 h in 0.1 M KCl aqueous electrolyte. Su and coworkers developed Fe–MoS2 nanosheets with atomic-level dispersion of Fe into MoS2.155 This Fe–MoS2 exhibits increased catalytic performance with a faster NRR process, thanks to the lower energy barrier (0.37 eV) of the potential-determining step and suppressed HER due to increased energy barrier (0.15 eV), compared to that of pristine MoS2 (0.50 eV, 0.03 eV). Consequently, the catalyst showed an NH3 production rate of 8.63 mgNH3 mgcat−1h−1 with a FE of 18.8%. He and colleagues introduced both Fe doping and S-vacancies in mixed-phase MoS2 to obtain VS-Fe-doped 1T/2H MoS2/C (VS-Fe–MoS2/C), which exhibited an ammonia yield rate of 17.8 ± 0.7 μg h−1 mg−1 and a FE of 9.2% under ambient conditions.156 The NRR performance of VS-Fe–MoS2/C outperforms those of Fe–MoS2/C and MoS2/C, showing the advantages of the multi-engineered synergistic approach in improving the NRR.


image file: d3mh00462g-f14.tif
Fig. 14 (a) Top panel: 3D topographic potential distribution of N2 adsorbed on Fe SAC (single-atom catalyst) protrusion immobilized on MoS2 (A1), pure MoS2 (A2), and Fe nanocluster immobilized on MoS2 (A3). Bottom panel: Calculated electric field EF (B1, C1, and D1), AFM images (B2, C2, and D2, scale bar of 50 nm), and experimentally measured EF distribution (B3, C3, and D3) of monolayer MoS2 (B1–B3), SACs-MoS2–Fe-1.0 (C1–C3), and SACs-MoS2–Fe-2.0 (D1–D3). (b) JFe/JMoS2 and IFe/IMoS2versus Fe loading amount. JFe and JMoS2 are the partial current densities of SACs-MoS2–Fe and MoS2, respectively, while IFe and IMoS2 are the EF intensities of SACs-MoS2–Fe and MoS2, respectively. (a and b) Reproduced with permission.154 Copyright 2020, Elsevier. (c) DFT simulation of Vs interacting with adsorbed dinitrogen molecules on undoped (top left) and Co-doped (bottom right) MoS2−x. The dinitrogen molecule binds to three Mo atoms near the S-vacancy in undoped MoS2−x and binds with one of the remaining Mo atoms near Vs in Co-doped MoS2−x. Bottom panel: Free energy profiles for the synthesis of NH3 from N2 and proton catalyzed by defective MoS2−x and Co-doped MoS2−x basal planes. (d) Optical images of CVD-grown MoS2−x and Co-doped MoS2−x triangular monolayers. The circular holes are the reaction windows on the basal planes for localized electrochemical measurements. The diameters of the holes are 4 μm. (e) Polarization curves of the catalyst in Ar and N2. The red and blue arrows indicate the regions assigned to the production of NH3 and H2, respectively. (c–e) Reproduced with permission.141 Copyright 2019, American Chemical Society.

For other metal atoms, Suryanto and coworkers reported a catalyst of Ru cluster loaded MoS2 with S-vacancies for NRR catalysis, delivering an NRR FE as high as 17.6% and a NH3 yield rate of 1.14 × 10−10 mol cm−2 s−1 at optimized temperature (50 °C) and potential (−150 mV).33 Compared to Ru/1T-MoS2, 2H-phase MoS2 decorated with amorphous Ru clusters (Ru/2H-MoS2) exhibited higher NH3 yields, also exceeding those of 2H-phase MoS2 and 1T-phase MoS2. The outstanding properties of Ru/2H-MoS2 in catalytic NRR stem from the synergistic interaction between Ru clusters as N2 binding sites and nearby isolated S-vacancies as H binding/provider sites for successive hydrogenations, thereby the interface between the Ru cluster and the S-vacancy is expected to play the key role in catalyzing NRR efficiently. Lou and coworkers discovered that the doped Co atoms collaborating with S-vacancies on MoS2 also facilitate the conversion of dinitrogen to ammonia under electrosynthesis conditions.141 The MoS2 monolayer shows very limited NRR activity due to the S passivation on both sides, whereas S-vacancy sites form a semi-open-edge configuration and thus result in the enhanced affinity of N2 molecules on the exposed three Mo atoms, which benefits from back-donation of electrons to dinitrogen's antibonding orbitals, facilitating the activation and subsequent dissociation of the N[triple bond, length as m-dash]N triple bond. Further Co doping makes N2 preferably to deviate from the tri-Mo center and to approach one of the Mo atoms, which leads to lower energy barriers for the adsorption and hydrogenation steps according to DFT simulations (Fig. 14c). The Co-doped MoS2−x catalyst prepared by a hydrothermal method showed a higher FE (over 10%) and yield rate (over 0.6 mmol h−1 g−1) than those of MoS2−x (1.7%, 0.32 mmol h−1 g−1), and the NRR activity is mainly from the basal plane according to the localized electrochemical measurements (Fig. 14d and e), indicating the effective modulation of the basal plane by Co doping and the S-vacancies.

The introduction of nonmetallic elements also contributes to the NRR activity. Fei and colleagues used a hydrothermal route to introduce P dopants to synthesize P-doped S-vacancy-rich MoS2.157 The introduced P atoms not only induce S-vacancies into MoS2, which is the active site for adsorption and conversion of N2 molecules, but also P itself modulates the electronic structure of MoS2, thus promoting the interfacial electron transfer between the catalyst and N2 molecules to enhance the adsorption of N2 molecules on the P-SV-MoS2 surface. The optimal catalyst had a superior electrocatalytic N2–NH3 conversion efficiency, with a FE value of 12.22% and an NH3 yield rate of 60.27 μg h−1 mg−1 at −0.6 V in 0.1 M Na2SO4. Chen and coworkers synthesized one-dimensional screw-like MoS2 nanosheets with oxygen partially replacing sulfur (1D-MoS2−xOy), delivering a remarkable ammonia yield rate of 5.56 × 10−8 mol s−1 cm−2 and a high faradaic efficiency of 9.86%.158 The enhanced performance is attributed to the partial atom substitution that reduces the Gibbs free energy of the potential-determining step.

Moreover, it is possible to optimize NRR performance by forming hybrids to lower the energy barriers or alter the NRR reaction pathways. Chen and colleagues constructed a 1T′-MoS2/Ti3C2 catalyst by a hydrothermal method.159 The hybrid catalyst attained a large NH3 yield rate of 31.96 μg h−1 mgcat−1 with a high faradaic efficiency of 30.75%. 1T′-MoS2/Ti3C2 presents a lower energy barrier than pure 1T′-phase MoS2 in the PDS step (*N2 → *N2H), indicating that the MoS2-based hybrid catalyst makes the activation and reduction of *N2 thermodynamically more favorable. Xu and coworkers developed a 1T-MoS2 NDs/g-C3N4 composite catalyst in which 1T-phase MoS2 nanodots were anchored on g-C3N4 and the composite catalyst showed the optimal reaction path of the NRR with both associative distal and alternating pathways.160 The 1T-MoS2 NDs/g-C3N4 catalyst achieved a large NH3 yield of 29.97 μg h−1 mgcat−1 and a high faradaic efficiency of 20.48% (see Table 7 for the comparison of NRR performance of the defective MoS2 catalysts). Zi and colleagues designed a S-vacancy-rich 1T-phase MoS2 monolayer confined on MoO3.161 In this hybrid, the appearance of depleted electrons in MoO3 and the aggregation of electrons in MoS2 indicate charge transfer from MoO3 to MoS2. The S-vacancy active sites strengthen the interaction between MoS2 and N2; specifically, the antibonding 2π* orbital of N2 is shifted towards the Fermi level due to the partial occupation of back-donated electrons from Mo d orbitals, resulting in a weakening of the highly inert N[triple bond, length as m-dash]N bond and thus an enhancement of the catalytic activity.

Table 7 Summary of NRR performance of defective MoS2-based catalysts
CatalystRef. Method Electrolyte Rate-determining step FE (%) NH3 yield rate J total (mA cm−2) Stability Potential (vs. RHE) Loading (mg cm−2)
MoS2/CC24 Hydrothermal growth 0.1 M Na2SO4 *N2 → *N2H 1.17 8.08 × 10−11 mol cm−2 s−1 ∼2.4 26 h −0.5 V N/A
MoS2/BCCF34 Hydrothermal method and annealing 0.1 M Li2SO4 *N2 → *N2H 9.81 43.4 μg h−1 mg−1 ∼1.3 12 h −0.2 V 10
Co-doped MoS2−x141 Hydrothermal process 0.01 M H2SO4 N2(g) → N2* Over 10 0.63 mmol h−1 g−1 ∼3.5 nA for monolayer N/A −0.3 V N/A
Defect-rich MoS2 nanoflower142 Hydrothermal synthesis 0.1 M Na2SO4 *NH2 → *NH3 8.34 29.28 μg h−1 mgcat−1 ∼0.4 20 h −0.4 V 0.4
Porous atomic layered 2H-MoS2143 One-step calcination 0.1 M HCl *N2 → *N2H 44.36 3405.55 μg h−1 mgcat−1 ∼0.1 36 h −0.1 N/A
Sub-monolayer MoS2−x145 Solid–solid reaction 0.05 M H2SO4 *NH2NH3 → *NH2 + NH3 24.7 17.2 μg h−1 mgcat−1 ∼0.13 20 h −0.2 0.25
1T/2H MoS2 hybrid146 Hydrothermal route 0.1 M Na2SO4 *N-*NH2 → *N-NH3 21.01 71.07 μg h−1 mgcat−1 ∼0.64 24 h −0.5 V 0.2
1T-MoS2-Ni144 Hydrothermal method 0.25 M LiClO4 *N2 → *N2H 27.66 1.05 μg min−1 cm−2 ∼2.8 13 h −0.3 V N/A
1T′′′-phase MoS2147 Deintercalation of potassium in KMoS2 0.1 M Na2SO4 *NH → *NH2 13.6 9.09 μg h−1 mg−1 ∼0.22 10 h −0.3 V 0.5
F-doped MoS2148 Hydrothermal method 0.05 M H2SO4 *N2 → *N2H 20.6 35.7 μg h-1 mgcat−1 ∼0.7 10 h −0.2 V 0.5
SACs-MoS2–Fe-2.0154 Hydrothermal grafting route 0.1 M KCl N[triple bond, length as m-dash]N → N = NH 31.6 ± 2 36.1 ± 3.6 mmol g−1 h−1 1.38 ± 0.12 50 h −0.2 V 0.15
Fe–MoS2155 Hydrothermal synthesis and mixture 0.5 M K2SO4 *N2 → *N2H 18.8 8.63 mgNH3mgcat−1 h−1 ∼0.18 N/A −0.3 V 0.8
VS-Fe-doped 1T/2H MoS2/C156 Hydrothermal reaction and annealing 0.1 M Na2SO4 N/A 9.2 17.8 ± 0.7 μg h−1 mg−1 ∼1 2 h −0.3/−0.5 V 0.17
P-doped S-vacancy-rich MoS2157 Hydrothermal route 0.1 M Na2SO4 *NH3 → NH3(g) 12.22 60.27 μg h−1 mgcat−1 ∼0.54 23 h −0.6 V 0.2
1D-MoS2−xOy158 Solvothermal method 0.1 M HCl *N2 → *N2H 9.86 5.56 × 10−8 mol s−1 cm−2 ∼13 12 h −0.35 V 0.7
1T′-MoS2/Ti3C2159 Hydrothermal method 0.1 M Na2SO4 *N2 → *N2H 30.75 31.96 μg h−1 mgcat−1 ∼13 12 h −0.7/−0.95 V 10 wt% of 1T′ phase
1T-MoS2 NDs/g-C3N4160 Hydrothermal and ultrasonic methods 0.1 M HCl *NHNH3 → *NH2 + NH3 20.48 29.97 μg h−1 mgcat−1 ∼0.5 24 h −0.3 V 10 wt% of 1T phase
SV-1T-MoS2@MoO3161 Hydrothermal method and calcination 0.05 M H2SO4 *N2 → *N2H 18.9 116.1 μg h−1 mgcat−1 ∼3.6 50 h −0.1/−0.2 V 0.36
Ru/2H-MoS2162 Redox reaction of LixMoS2 and RuCl3 0.01 M HCl *NH2 + *NH2 + *H → *NH3 + *NH2 17.6 1.14 × 10−10 mol cm−2 s−1 ∼0.30 4 h −0.15 V 1


However, it is worth noting that although the NRR reaction has achieved high FE based on several strategies mentioned above, the ammonia yield rate remains very limited due to the dominance of the HER in the electrochemical process, especially at higher overpotential, and the selectivity of the NRR continues to be an issue for MoS2-based electrocatalysts. Therefore, there is an urgent need for more attempts on MoS2 guided by structural defects, crystal phase regulation, S-site interactions, atomic doping, hybrid design, and other engineering approaches to further improve the NRR performance by enhancing N2 adsorption, reducing the activation energy, and suppressing the HER process.

4.2. Carbon dioxide reduction

The overconcentration of CO2 accumulated from excessive consumption of fossil fuels has become one of the major contributors to global warming. Replacing fossil fuels with clean energy sources, such as photovoltaic and green hydrogen to prevent carbon emissions and developing CO2 capture/storage/conversion technologies to reduce the current CO2 concentration are considered as promising technologies to address the environmental issues. In particular, the conversion of CO2 into valuable chemicals through direct electrochemical reduction powered by renewable energy sources such as solar, wind, and tidal energy can provide both environmental and economic benefits. Electrocatalysts including Au, Ag, Pd, Zn, Sn, Bi, Cu, carbon materials, metal oxides, etc.163 have demonstrated their catalytic performance in CO2 electroreduction, although they still face different challenges such as high cost (for precious metals), high overpotential (for some metals), low-value products (mainly the C1 product), low faradaic efficiency, and low selectivity to target products, which hinder their implementation at the commercial scale. To this end, MoS2 having shown superior catalytic performance in the CO2 reduction reaction offers a promising alternative.

Francis and coworkers disclosed that MoS2 single crystal terraces or films with low edge-site densities could convert CO2 into 1-propanol in aqueous electrolytes with FE values of ∼3.5% or ∼1%, respectively.167 In contrast, the molybdenum-terminated edges of MoS2 favor the formation of carbon monoxide (CO) in an ionic liquid. Asadi and coworkers found that the MoS2 bulk with a Mo-terminated edge, compared to Ag, exhibited a much higher current density of 65 mA cm−2 at −0.764 V with a FE value close to 98%, when used as the catalyst for the reduction from CO2 to CO in 4 mol% EMIM-BF4 aqueous solution. Such a superior performance is attributed to the much higher d-electron density in the vicinity of the Fermi level of Mo-edge atoms so that the Mo d-electron can form metallic edge states and freely supply excess d electrons to the attached reactants at the Mo edge (Fig. 15a).23 Accordingly, a vertically aligned MoS2 nanosheet terminated by Mo atoms was also fabricated, further achieving a 2-fold higher current density for the CO2 reduction than the bulk state (Fig. 15b). With regard to the CO2–CO pathway (* + CO2 + H+ + e → COOH*; COOH* + H+ + e → CO* + H2O; CO* → CO + *), the formation of COOH* is endergonic and thus known as the rate-limiting step on certain metal surfaces (e.g., Pd, Au, Cu, and Ag), whereas it is exergonic on MoS2 metal edge sites because of the strong binding of the adsorbed intermediate to the edge thanks to its d-band center close to the Fermi level. In addition, more stable CO* can be formed on the metallic edge, which is also an exergonic process that leads to a lower overpotential for CO2 conversion. However, the desorption of CO becomes a rate-limiting step since the strong binding of the adsorbed CO on MoS2 (Fig. 15c) suppresses its release.164,165 Therefore, accelerating the desorption of CO product from the MoS2 catalyst surface can facilitate CO2 conversion with a higher turnover frequency and FE. Various defects engineering methods on MoS2, including doping, alloying, hybridization, and surface decoration, have been proven to be effective in stimulating the CO desorption process, which will be discussed sequentially in the next section and finally we will discuss defect strategies that facilitate the production of high-value gas/liquid products by electrocatalytic or photo(electro)catalytic means.


image file: d3mh00462g-f15.tif
Fig. 15 (a) PDOS for the spin-up channels of: the Mo atom at the edge and the Mo atom in the lattice (left panel); s, p, and d orbitals of the Mo-edge atom (middle panel); PDOS of d band of the Mo-edge atom, the Ag atom from the bulk and Ag-slab (right panel). (b) Polarization curves for CO2 reduction of bulk MoS2 and VA (vertically aligned) MoS2. The inset shows partially enlarged curves. (a and b) Reproduced with permission.23 Copyright 2014, Springer Nature. (c) DFT calculated free energy profiles for CO2 electroreduction to CO on Ag(111), Ag55 nanoparticles, MoS2 (in black line), WS2, MoSe2, and WSe2 nanoflakes. Reproduced with permission.164 Copyright 2016, AAAS. (d) Reaction pathways of CO2 → CO on MoS2, NbS2, MoS2-Nb-1 (replacing the second row of Mo atoms from the edge with Nb atoms), and MoS2-Nb-2 (replacing the third row of Mo atoms from the edge with Nb atoms). (e) CO formation TOF for VA-Mo0.95Nb0.05S2, pristine VA-MoS2, and Ag nanoparticles at different applied overpotentials. (d and e) Reproduced with permission.165 Copyright 2016, American Chemical Society. (f) Calculated free energy diagrams for CO2 electroreduction to CO at MoS2 and N–MoS2 electrodes. Reproduced with permission.166 Copyright 2019, Elsevier.

Abbasi and coworkers achieved decreased binding strength between the Mo edge and CO by substitutional doping of Nb atoms into the MoS2 structure, accelerating the turnover of CO desorption, while retaining the exergonic behavior of forming COOH* and CO* (Fig. 15d).165 As a result, in the overpotential range of 100–300 mV, the CO formation TOF of the p-type doped vertically aligned VA-Mo0.95Nb0.05S2 with 5% Nb concentration was about 1 and 2 orders of magnitude higher than that of pristine VA-MoS2 and Ag nanoparticles, respectively (Fig. 15e). For other metals, Mao and colleagues systematically investigated the electrocatalytic activity of the CO2RR on the MoS2 edge doped with transition metal atoms using high-throughput DFT calculations.168 It was found that V, Zr and Hf dopants in MoS2 could significantly reduce the binding energies of CO, thus promoting the desorption of CO from the MoS2 edge and achieving optimal CO2RR performance. Beyond metal doping, Lv and coworkers suggested that non-metallic nitrogen doping also helps to weaken the adsorption strength of CO* on the Mo atoms of N-MoS2, where N doping lowered the CO desorption Gibbs free energy (ΔG) from 0.53 eV (undoped MoS2) to 0.49 eV (Fig. 15f), leading to greatly enhanced catalytic activity towards the CO2RR.166 Huang and coworkers combined experimental and theoretical studies to elucidate that compared to single N or P doping, N,P co-doped MoS2 achieves higher current density and FE and synergistically reduces the binding strength between the Mo sites and CO moieties, resulting in improved performance.169

In addition to the doping strategy, Xu and coworkers developed an alloyed MoSeS monolayer to weaken its binding strength with CO.170 Unlike MoS2 and MoSe2 where the charge density is mainly distributed on Mo atoms, MoSeS has the shortened Mo–S bond and the lengthened Mo–Se bond, leading to the off-center charge density of the Mo atoms (Fig. 16a). Consequently, the overlap of the d orbitals of Mo atoms and the p orbitals of C atoms in CO* is reduced and thus weakens the CO adsorption strength as verified by the elongated C–Mo bond and facilitates the rate-limiting CO desorption step, as evidenced by the lowest CO desorption temperature/potential on the MoSeS alloy in the experiment. The off-center charge density of the Mo atom also favors the stabilization of the COOH* intermediate on the charge accumulation region between two neighboring Mo atoms, which leads to a higher exothermic and spontaneous COOH* formation process. As a result, the higher FE than MoS2 (16.6%) and MoSe2 (30.5%) was obtained for the MoSeS monolayer (45.2%) at −1.15 V, with a current density of 43 mA cm−2, which was about 2.7 and 1.3 times higher than those of MoS2 and MoSe2, respectively.


image file: d3mh00462g-f16.tif
Fig. 16 (a) Charge density distributions of conduction band edges (left panel) and crystal structures (right panel) of the MoSeS alloy monolayer (row 1), MoS2 monolayer (row 2), and MoSe2 monolayer (row 3). Reproduced with permission.170 Copyright 2017, John Wiley & Sons Inc. (b) Left panel: Free energy profiles for CO2RR catalyzed by MoS2 basal plane, NC (N-doped carbon), MoS2 edge@NC, and MoS2 edge. Right panel: Illustration of electron density changes in the NCMSH model, where the turquoise color regions stand for holes upon contact between the MoS2 edge and NC. Reproduced with permission.171 Copyright 2019, John Wiley & Sons Inc. (c) A thin film of H-E-MoS2 with wettability regulation is immobilized on the GC electrode substrate, appearing hydrophobic and aerophilic. The TPCP (liquid/gas/solid tri-phase contact point) permits more contact of the catalyst with CO2. (d) CO and H2 FE% for the H-E-MoS2 catalyst at different applied potentials, and the inset shows water droplets on the surface of H-E-MoS2, suggesting its hydrophobicity. (c and d) Reproduced with permission.172 Copyright 2018, John Wiley & Sons Inc.

Li and coworkers proposed a hybridization strategy of edge-exposed 2H-phase MoS2 with N-doped carbon (NCMSH).171 Attributed to the electron transfer from the electron-rich N-doped carbon to the electron-deficient Mo edge by theoretical calculations, the acquired electrons around the Mo atom are believed to facilitate the potential limiting step (see Fig. 16b for free energy profiles of the CO2RR and the electron density change in the NCMSH model), i.e., the reduction of COOH* and the recovery of the original Mo site. This catalyst with a large exposed MoS2 edge at the N-doped carbon site displayed a FE of 92.68% and a high CO production rate of 31.80 mA cm−2, which corresponds to a threefold enhancement of the MoS2 without N-doped carbon. Moreover, Lv and coworkers devised a surface modification method to decorate the exfoliated MoS2 with fluorosilane (FAS).172 On the basis of changing the electronic properties of the edge Mo atoms, the decorated FAS on the MoS2 surface brings down the Gibbs free energy ΔG (0.72 eV) of the rate-limiting CO* desorption step from the undecorated MoS2 (1.32 eV), thereby weakening the adsorption strength of CO with Mo atoms and accelerating the desorption of CO product. Furthermore, the hydrophobic and aerophilic properties of the FAS-decorated MoS2 surface suppress the HER process and create sufficient three-phase contact points for the CO2 reaction (Fig. 16c). As a result, the hydrophobic exfoliated MoS2 (H-E-MoS2) obtained a higher CO FE% (81.2%) (see Fig. 16d for CO and H2 FE% of H-E-MoS2) and a lower onset potential (−0.24 V) than those of pure exfoliated MoS2 (41.2%; −0.3 V) and bulk MoS2 (19.8%; −0.43 V). Li and colleagues synthesized amorphous MoSx hybrids on polyethylenimine (PEI)-modified reduced graphene oxide (rGO-PEI-MoSx), in which PEI also suppressed HER and stabilized the CO2˙ intermediate species during CO2 reduction, thus synergistically improving the catalytic activity of MoSx, attaining an onset overpotential of 140 mV and a maximum FE of 85.1% (see Table 8 for the comparison of CO2RR performance of the aforementioned MoSx catalysts).173

Table 8 Summary of CO2RR performance of defective MoSx-based catalysts
CatalystRef. Method Electrolyte Main product Rate-determining step Potential (vs. RHE) FE (%) Partial current density (mA cm−2) Stability (h)
Bulk MoS2 with Mo-terminated edge23 Natural layered bulk MoS2 4 mol% EMIM-BF4/ H2O CO N/A −0.764 V ∼98 ∼63.7 10
MoS2 single-crystal terrace167 Mechanical exfoliation 0.10 M Na2CO3 1-propanol N/A −0.59 V ∼3.5 ∼0.26 10
VA-Mo0.95Nb0.05S2165 Chemical vapor deposition 50 vol% EMIM-BF4/ H2O CO CO* → CO + * −0.8 V 82 ∼194 N/A
N-MoS2@NCDs-180166 Solvothermal method 6 mol% EMIM-BF4/ H2O CO CO* → CO(g) −0.9 V 90.2 ∼33 10
N,P-MoS2 NSA169 Hydrothermal growth and NH3 treatment 6 mol% EMIM-BF4/ H2O CO CO* → CO + * −0.9 V 91.5 ∼7 30
MoSeS monolayer170 Liquid–liquid interface-mediated strategy 4 mol% EMIM-BF4/ H2O CO CO* → CO + * −1.15 V 45.2 ∼19 10
Edge-exposed 2H-phase MoS2 hybridized with NC171 Solvothermal method and pyrolysis 4 mol% EMIM-BF4/ H2O CO COOH* + H+ + e → *+ CO + H2O −0.7 V 92.68 31.80 24
Hydrophobic exfoliated MoS2172 Ball milling and CVD reaction with FAS 6 mol% EMIM-BF4/ H2O CO CO* → CO + * −0.9 V 81.2 ∼18 10
rGo-PEI-MoSx173 Electrodeposition 0.5 M NaHCO3 CO N/A −0.65 V 85.1 4 16


Besides the pursuit of higher FE and lower overpotentials of CO formation on defective MoS2 electrocatalysts, liquid products are also desirable. Although MoS2 single crystals have shown success in the production of liquid 1-propanol, the low FE (∼3.5%) hinders its practical use for the CO2RR. Engineering MoS2 through defective strategies can facilitate the production of high-value gas/liquid products by electrocatalytic or photo(electro)catalytic means. From the perspective of electrocatalysis, Ru@1T′-MoS2, Fe@MoS2, Co@MoS2, Ni@MoS2, MoS2-IrCl3 and monolayer MoS2 with 2S-vacancies are potential catalysts for the production of CH4 in different theoretical works.174–177 For the CH3OH production, Pt@1T′-MoS2 and MoS2-IrF3 are promising.174,176 For the generation of HCOOH, Re@1T′-MoS2 and Zn@1T′-MoS2 are achievable.174 For the 1-propanol production, VS-MoS2 is hopeful.178 These vacancies or doping defects alter the catalytic free energy of the intermediate species by breaking the scaling relationship, changing the orbital orientation, or exploiting steric hindrance.176,177,179 For example, Nørskov and coworkers proposed that in transition metal-doped MoS2, there are two scaling relationships: one for the doped metal and one for sulfur.179 The metal site favors CO* adsorption, while the S site favors COOH* and CHO*, so that during the reduction of CO* to CHO*, the intermediate switches from the metal site to S site, thus breaking the constraint of scaling relation and leading to the stabilization of CHO* species and the formation of hydrocarbons. In another Jiang's report, the authors proposed a bowl site structure, which can be a 2S-vacancy or a Mo atom substituting a S2 column and the bowl sites can achieve the selective reduction of CO2 to CH4.177,180 In this configuration, *CO formed by CO2 reduction is adsorbed in the bowl site and due to the steric hindrance, the adsorbate on the bowl site forms repulsive interaction with external species, thus preventing the hydrogenation of carbon and enabling the hydrogenation of oxygen, resulting in the formation of *COH instead of *CHO. Similarly, in the next step, *COH has difficulty forming *CHOH, but instead dehydrates to form *C, thus providing the basis for subsequent protonation to form CH4. In contrast, on the molybdenum sulfide surface without a bowl structure as well as steric hindrance, multiple catalytic pathways coexist, leading to a diversity of products. On the basis of the above theoretical work, relevant experimental evidence is expected to emerge in the near future.

The band gap of MoS2 (1.17 eV) cannot match the CO2 reduction potential181 and the photocatalytic activity of CO2 could be promoted by doping or constructing a heterostructure to optimize the energy band structure or facilitate carrier transfer. Peng and coworkers developed Co-doped MoS2 with an optimized valence band (0.89 V) and conduction band (−0.52 V) to match the reduction potential of CO2 and the oxidation potential of H2O, thus obtaining photocatalytic activity.182 The methanol yield rate in the photoelectrochemical (PEC) reactor reached 35 mmol L−1. Wang and colleagues reported a Z-scheme heterojunction constructed of 3D-SiC@2D-MoS2.183 Thanks to the more positive valence band and higher hole mobility of MoS2, and the more negative conduction band and higher electron mobility of SiC, the material achieved CO2 to CH4 conversion without sacrificial reagents under visible light (λ ≥ 420 nm) irradiation, delivering a CH4 yield rate of 323 μL g−1 h−1 and stable operation for 40 h. Yu and coworkers synthesized a d-UiO-66/MoS2 by integrating MoS2 nanosheets into porous defective UiO-66 for photocatalytic conversion of CO2 and H2O to CH3COOH under visible light irradiation without adducts.184 In this system, MoS2 contributes to the reduction of the band gap (Eg), making it possible to generate active carriers under visible light; MoS2 also facilitates the adsorption of CO2. Moreover, bridging Mo–O-Zr on the interface promotes the transfer of photogenerated carriers and lowers the energy barrier for C–C coupling. As a result, the evolution rate and selectivity of CH3COOH obtained by d-UiO-66/MoS2 reached 39.0 μmol g−1 h−1 and 94%, respectively. Overall, building on the above success, more types of liquid products and higher yields remain to be expected.

4.3. Metal–sulfur batteries

The current energy density of commercial lithium-ion batteries (≈300 mA h g−1)185 is approaching its theoretical limits, so there is an urgent need to develop next-generation energy storage batteries with high energy density, fast charging, long cycle life, environmental friendliness, and excellent cost-effectiveness to meet the growing demand for industrial electrification. Among the advanced alternatives, metal–sulfur batteries are promising candidates benefiting from their high theoretical capacity of sulfur (1675 mA h g−1),186 low-cost and abundant sulfur resources, and environmental benignity. Lithium–sulfur (Li–S) batteries with lithium as the anode and S as the cathode have a theoretical specific energy density of 2600 W h kg−1,187 while the sodium–sulfur (Na–S) batteries with Na as the anode possess an energy density of 1274 W h kg−1,36 making both of them promising candidates. Upon discharging of Li–S batteries (LSBs), the lithium at the anode is oxidized to form Li+ ions (Li → Li+ + e) that move towards the cathode through a diaphragm. In contrast, the sulfur at the cathode is gradually reduced from cyclic octasulfur (S8) to soluble long-chain lithium polysulfide (LiPS) intermediates (Li2Sn, 3 ≤ n ≤ 8) that undergo further reduction to form insoluble short-chain Li2S2 and Li2S (S8 + 16Li++ 16e → 8Li2S) (Fig. 17a).35 On the other hand, the subsequent charging process undergoes a reverse reaction that reconverts Li2S to elemental sulfur (8Li2S → S8 + 16Li+ + 16e) and lithium metal (Li+ + e → Li), resulting in a reversible discharge/charge cycle. Several obstacles, including poor conductivity of S, the shuttle effect of LiPSs, volumetric expansion, sluggish kinetics of sulfur transformation, self-discharge, and lithium dendrite, restrict the commercial applications of LSBs.35,188,189 As one of the key issues, the shuttle effect arises from the diffusion of intermediate LiPSs from the cathode to the anode, which leads to irreversible loss of sulfur-active material, passivation of the anode, and increased internal resistance due to their high solubility in the organic electrolytes, resulting in low Coulombic efficiency, decreased battery capacity, and poor cycling stability.190,191
image file: d3mh00462g-f17.tif
Fig. 17 (a) A typical charge/discharge profile for a Li–S battery, where the molecular model is labeled at the corresponding stage. Reproduced with permission.35 Copyright 2017, John Wiley & Sons Inc. (b) Cyclic voltammograms of symmetric cells with identical MoS2−x/rGO electrodes in electrolytes with and without Li2S6. Four distinct peaks are marked, showing high reversibility. (c) Comparison of rate performance of the MoS2−x/rGO/S, MoS2/rGO/S and rGO/S cathodes at different C-rates. (b and c) Reproduced with permission.198 Copyright 2017, Royal Society of Chemistry. (d) Adsorption energies of Li2S6 and Li2S8 on LE-MoS2 (top) and conversional MoS2 (bottom). Carbon interlayers between MoS2 slabs are shown. Layer spacing between MoS2 monolayers is indicated and the adsorption energies are labeled. In situ sulfur K-edge XANES of the sulfur-loaded S/LE-MoS2 electrode at (e) discharge and (f) charge status. The sulfur peak (2472.0 eV) gradually diminishes, while LiPS peaks (2470.1 eV) appear and reach a maximum when discharged to ∼2.1 V. Further discharge leads to the characteristic peaks of 2473.0 and 2475.3 eV for Li2S, while no sulfur or LiPS (as shown by the cross) peaks are observed at 1.7 V. When charged back to 3.0 V, the sulfur peak reappears, accompanied by the disappearance of the Li2S (as shown by the cross). (d–f) Reproduced with permission.199 Copyright 2020, American Chemical Society.

The shuttle effect can be suppressed by reducing the solubility of LiPSs in the electrolyte through physical blocking or chemisorption strategies and by accelerating the conversion of LiPSs to solid Li2S2/Li2S using catalysts.191,192 Similar to the reaction mechanism of LSBs, room-temperature sodium–sulfur batteries (Na–S batteries) face the same shuttle effect and thus also require the adsorption and rapid conversion of soluble sodium polysulfides (NaPSs).193 Among the polysulfide conversion catalysts, MoS2 is considered as one of the powerful contenders, thanks to its strong adsorption to polysulfides (i.e., LiPSs, NaPSs) through chemical binding and its intrinsic high chemical activity.194,195 MoS2 edge sites with unsaturated coordination exhibit stronger interaction and catalytic activity, while the basal plane with saturated coordination is relatively inert in catalyzing polysulfide conversion,187,196,197 similar to HER catalysis. Optimization of MoS2 catalysts using defect engineering provides a route to suppress the shuttle effect and to improve the performance of metal–sulfur batteries. As for two-dimensional materials, process optimization can be performed by creating structural defects, vacancies, doping and alloying within the layers, by interlayer expansion and interlayer ion intercalation, and by constructing heterojunctions or hybridization outside the layers, which will be discussed in turn in the next section. In addition, for non-2D MoSx, excellent catalytic properties can be obtained through methods such as amorphization, which will be discussed at the end.

For the within-layer engineering approach, Lin and coworkers reported a sulfur-deficient 2H-phase MoS2 nanoflake electrocatalyst via liquid phase exfoliation and the removal of sulfur atoms using thermal treatment in hydrogen.198 The catalytic behavior of MoS2−x for polysulfide redox reactions was evaluated in Li2S6 electrolyte, where MoS2−x/rGO exhibited high activity and reversibility with four distinct peaks implying two-step lithiation of sulfur to Li2S and its reversible reaction (Fig. 17b), while MoS2/rGO exhibited broad redox peaks with remnants. In terms of practical performance, the MoS2−x/rGO/S composite used as the cathode in coin cells with a lithium metal anode delivered an increased capacity of 826.5 mA h g−1 compared to those of MoS2/rGO/S (473.3 mA h g−1) and rGO/S (161.1 mA h g−1) at a rate of 8 C (Fig. 17c) and enhanced cycle stability of the sulfur cathode with a capacity decay of 0.083% per cycle over 600 cycles at a 0.5 C rate. The improved rate performance is attributed to the higher affinity of MoS2−x/rGO and the accelerated conversion kinetics of polysulfides on the sulfur deficiency sites that are regarded as catalytic centers, leading to a reduced sulfide loss and therefore a sustained cycling capacity. Du and colleagues fabricated 1T-phase MoS2 (∼90%) monolayers by converting Mo-based MXenes (Mo2CTx and Mo1.33CTx) in H2S at high temperatures.200 The Mo vacancy in Mxenes promotes the gliding of the sulfur layer to form 1T-phase MoS2, while Mo1.33CTx-derived MoS2 with more Mo vacancies exhibits intense chemisorption and high catalytic activity for LiPSs, delivering a reversible capacity of 736 mA h g−1, an outstanding rate capability of 532 mA h g−1, and good stability for 200 cycles in Li–S batteries.

LIB performance can also be improved when S-vacancy defects and phase tuning are combined. Zhang and coworkers developed an MXene/1T-2H MoS2-C catalyst with plentiful positively charged S-vacancies and abundant 1T phase, which was verified to exhibit stronger trapping and adsorption ability of polysulfides (including S8/S62− and S42−/S62−) than MXene/2H MoS2-C and MXene, through visual observation (i.e., color changes of a Li2Sx solution before and after adsorption) and UV-vis spectroscopy during discharge.201 As a result, MXene/1T-2H MoS2-C achieved a remarkable improvement in battery performance, with a capacity of 1194.7 mA h g−1 at 0.1 C and a fading rate of 0.07% per cycle over 300 cycles at a 0.5 C rate. Moreover, the MXene/1T-2H MoS2-C cathode showed reliable performance in soft-package batteries, even after the bending test. Cheng and coworkers suppressed the shuttle effect by using sulfur-deficient 1T-phase MoS2 nanoflower-decorated graphene (FM@G) as both the cathode and separator, where the S-deficient metallic MoS2 with an accessible catalytic surface at the cathode boosts chemisorption and catalytic conversion of polysulfides, while the separator further suppresses the permeation of LiPSs to the anode.202 The synergistic design resulted in excellent cyclability with a decay rate of 0.058% per cycle for 500 cycles at 1.0 C rate and an initial capacity of 1057 mA h g−1. More similar catalyst systems incorporating defects and phase modulation, such as edge-rich 1T-phase MoS2 nanodots203 and CF@2H/1T MoS2,204 have also been shown to be effective in the adsorption and catalysis of LiPSs.

Heteroatom doping within the MoS2 layer improves the adsorption capacity and conversion kinetics of polysulfides by inducing transformation into the metallic 1T phase, improving the electrical conductivity, and introducing S-vacancies. Liu and coauthors recently pointed out that the formation energies of the 1T phase and S-vacancy after Co doping are both lower than those of the 2H phase by DFT calculations, and the decomposition energies of Li2S4, Li2S2, and Li2S on the Co-doped 1T-phase MoS2 are lower than those of 1T-phase MoS2.205 In the experiments, a hydrothermal method and the use of excess Co precursor were adopted to synthesize Co-doped MoS2, where Co was substitutionally doped. In such a catalyst, abundant S-vacancies were introduced and the 1T content was increased. As a result, the Co–MoS2 on graphene showed stronger static adsorption of Li2S6, more distinct redox peaks in the symmetric tests, a lower fading rate of 0.029% per cycle over 1000 cycles at the 1 C rate, and a higher capacity of 941 mA h g−1 at 2 C rate in Li–S coin cells. Lin and coworkers employed a cobalt and phosphorus co-doping strategy, where cobalt induced the phase change to metallic 1T phase and phosphorus had an affinity to cobalt, to form Co–P coordinated sites with high catalytic activity for polysulfide conversion.187 Consequently, the sulfur cathode using the Co–P co-doped MoS2 catalyst (P-Mo0.9Co0.1S2-2) exhibited a higher rate capability of 931 mA h g−1 at the 6 C rate than those of the cathodes using Mo0.9Co0.1S2 (633 mA h g−1) and MoS2 (338 mA h g−1) catalysts. Moreover, the P-Mo0.9Co0.1S2-2/S cathode provided a capacity decay rate of 0.046% per cycle over 600 cycles at 1 C rate, implying the importance of a dopant in MoS2 catalysts for polysulfide conversion. Similar to the doping of cobalt and phosphorus, Li and coworkers introduced nickel and phosphorus, which also induced the phase transition of MoS2 from the 2H phase to 1T phase and facilitated the formation of Ni–P coordination, which improves the rate performance and cycling performance by accelerating the adsorption and conversion of polysulfides, as well as suppressing the shuttle effect.206

In addition to the doping strategy, alloying approaches have also shown great effectiveness in Li–S batteries. Bhoyate and coauthors developed a mixed phase 2D Mo0.5W0.5S2 (2H-1T phases) alloy, fabricated by co-sputtering and sulfurization methods, which was verified to demonstrate strong adsorption to Li2S6.207 Compared with MoS2-CNT-S and WS2-CNT-S, the Mo0.5W0.5S2-CNT-S cathode achieved low interfacial and charge-transfer resistances, a high specific capacity of 1228 mA h g−1 at 0.1 C, and high cycling stability. Bhoyate and colleagues synthesized a mixed 2H/1T phase MoxW1−xS2−y alloy (D-MoWS) with tunable defects, which demonstrated a high specific capacity of 1586 mA h g−1 and an excellent areal capacity of 13.5 mA h cm−2 at high sulfur loading.208 A mechanistic study has proven that the D-MoWS catalyst accelerates the LiPS conversion, and prevents the shuttling effect and Li–metal corrosion.

As for the interlayer engineering approach, Luo and coworkers noted that the interlayer expansion of MoS2−x nanosheets with S-vacancies enhances the immobilization and catalysis of NaPSs in Na–S batteries, where the S-vacancy increases the electrical conductivity and binding strength by altering the distribution of valence electrons, and the interlayer-expanded structure offers additional active sites that may enhance the adsorption efficiency, leading to strong catalytic conversion towards NaPSs.193 More stable interlayer structure can be obtained by intercalation methods, such as carbon, Li+, and Na+ intercalation. Pan and colleagues fabricated layer-spacing-enlarged MoS2 (LE-MoS2) by carbon intercalation.199 DFT simulations revealed that compared to the unexpanded MoS2, the intercalated MoS2 possesses higher adsorption energies for Li2S6 and Li2S8 (Fig. 17d), stronger electronic coupling between LiPSs and LE-MoS2, and lower cleavage energies from Li2S4 to Li2S, resulting in stronger chemisorption and better electrocatalytic performance. When used as the electrode during discharge/charge of a Li–S cell, the sulfur-loaded LE-MoS2 exhibited complete conversion between sulfur and Li2S (i.e., lithiation of sulfur and reversible oxidation of Li2S to LiPSs/sulfur) as confirmed by sulfur K-edge XANES (Fig. 17e and f); and no obvious color change from soluble Li2S6 was observed after disassembling the cell, indicating its outstanding catalytic effect and adsorption to LiPSs. For these reasons, LE-MoS2 showed a high initial capacity of 1550 mA h g−1 and a low decay rate (0.06% per cycle) over 500 cycles at the 1 C rate. Liu and coworkers developed a Li+ intercalated MoS2 by in situ electrochemical methods based on the discharged plateaus, where a cut-off voltage below 1.1 V facilitated the Li+ insertion process to enlarge the interlayer space and induce phase transition from 2H to 1T phase (Fig. 18a), which was stable and irreversible even undergoing the next charging process.209 According to CV and electrochemical impedance spectroscopy (EIS) analyses, the TiN/1T-MoS2/S cathode exhibited faster charge-transfer and redox kinetics and thus delivered a better rate performance of 689 mA h g−1 at 2 C (Fig. 18b) and a lower capacity fading rate of 0.051% per cycle at 1C for 800 cycles than those of the TiN/2H-MoS2/S cathode. Similar to Li+ intercalation, Wang and coworkers proposed a Na+ intercalation strategy, which was also realized by discharging to a cut-off voltage of 0.8 V in a room-temperature Na–S battery with S/MoS2 as the starting cathode (Fig. 18c).36 Compared to the S/MoS2 electrode (cut-off voltage of 1.0–2.8 V), the S/NaxMoS2 electrode (cut-off voltage of 0.8–2.8 V) showed better rate capacity (Fig. 18d), less polarization, smaller charge transfer resistance, higher capacity (774.2 mA h g−1) after 800 cycles, and better retention with a fade rate of 0.0055% per cycle for 2800 cycles. The stronger adsorption energies of Na2S2 and Na2S on NaxMoS2 contribute to the adsorption of polysulfides and the nucleation of Na2S, which make the conversion of polysulfides more reversible and thus result in higher cyclability in batteries.


image file: d3mh00462g-f18.tif
Fig. 18 (a) Left panel: CV of batteries with Li metal as the anode and TiN/2H-MoS2/S as the cathode, the peaks located around 2.3, 2.0 and 1.1 V can be ascribed to the transformation from long-chain LiPSs to short-chain LiPSs, from short-chain LiPSs to Li2S2 and Li2S, and phase transformation from 2H-phase MoS2 to 1T-phase MoS2. Right: Galvanostatic discharge–charge profiles in 2.8–1.0 V, the plateau located at 1.1 V indicates the facile conversion from the 2H phase to 1T metallic phase. (b) Rate performance of TiN/1T-MoS2/S and TiN/2H-MoS2/S. (a and b) Reproduced with permission.209 Copyright 2021, John Wiley & Sons Inc. (c) Schematic of the discharge curve of the electrode and the molecular conversion catalyzed by MoS2. MoS2 can catalyze the conversion of S8 to Na2S6 (phase I) and of long-chain Na2S6 to short-chain Na2S2 (phase II); the sodiated NaxMoS2 after discharge to 0.8 V promotes the conversion of Na2S2 to Na2S (phase III). (d) Rate capability of the S/MoS2/NCS composite. The deeper discharge to 0.8 V leads to more rate capability due to the formation of sodiated NaxMoS2, where Mo4+ has been partially reduced. (c and d) Reproduced with permission.36 Copyright 2021, John Wiley & Sons Inc. (e) First charge/discharge curves of Co9S8@MoS2/CNF, Co9S8/CNF, MoS2/CNF, CNFs (carbon nanofibers), and without interlayer. The voltage hysteresis (ΔH) is labeled. Reproduced with permission.191 Copyright 2020, American Chemical Society.

Establishing heterojunctions or hybridization outside the layers is also an effective strategy to suppress the shuttle effect. Li and coworkers reported a Co9S8@MoS2 core–shell heterostructure in which the conductivity of MoS2 was improved by introducing Co9S8 that brings new states near the Fermi level.191 Furthermore, the adsorption of LiPSs and the conversion of LiPSs to Li2S can be improved due to stronger adsorption and dissociation energies as revealed by theoretical calculations. Compared with the cathodes of Co9S8/CNF and MoS2/CNF, the Co9S8@MoS2 heterostructure-embedded carbon nanofibers (Co9S8@MoS2/CNF) showed a stronger affinity to LiPSs, higher discharge capacity (1106 mA h g−1 after 100 cycles at 0.5 C), and lower polarization in LSBs (Fig. 18e). Moreover, the Co9S8@MoS2/CNF achieved superb cycling stability with a reversible capacity of 794 mA h g−1 at 1 C and a fade rate of 0.091% per cycle over 400 cycles, as well as high capacities at high sulfur loadings. Similar heterostructure catalysts, such as MoS2−x–Co9S8−y with Mo–S–Co heterojunction210 and Fe7S8–MoS2,211 also perform well in the catalytic conversion of LiPSs. Moreover, Wang and colleagues used a MoS2–MoN heterostructure to accelerate the conversion of polysulfides.212 MoS2 has moderate adsorption energy to polysulfides and a low lithium-ion diffusion barrier, while MoN provides coupling electrons through a redox reaction and has high electrical conductivity, thus MoS2–MoN provides better adsorption and regulation of long-chain LiPSs, stronger lithium-ion diffusion capability, and faster reaction kinetics. Along the enhanced adsorption–diffusion–conversion process based on MoS2–MoN/S cathode, 1000 cycles were achieved at 1 C with a decay rate of as low as 0.039% per cycle (see Table 9 for a more comprehensive comparison of defective MoS2 catalysts).

Table 9 Summary of metal–sulfur battery performance of defective MoS2-based catalysts
CatalystRef. Method Metal–sulfur battery Sulfur loading (mg cm−2) Rate performance Initial discharge capacity Cycle Capacity retention Capacity decay rate (%) Coulombic efficiency (%)
S/MoS2/NCS36 Carbonization and sulfurization Na–S N/A 470.7 mA h g−1 @5 A g−1 ∼864 mA h g−1 @0.2 A g−1 800 cycles@0.2 A g−1 774.2 mAh@0.2 A g−1 0.013 ∼100
CC/1T-MoS2186 Hydrothermal method Li–S 2.1 893 mA h g−1@2 C 1233 mA h g−1 @0.5 C 500 cycles@0.5 C 919 mA h g−1 @0.5 C 0.051 87.1
P-Mo0.9Co0.1S2187 Hydrothermal method Li–S 2 931 mA h g−1@6 C 1233 mA h g−1 @1 C 600 cycles@1 C 886 mA h g−1@1 C 0.046 N/A
Co9S8@MoS2/CNF191 Electrospinning and sulfuration Li–S 3 477 mA h g−1@6 C 1250 mA h g−1 @1 C 400 cycles@1 C 794 mA h g−1@1 C 0.091 ∼100
FL-MoS2−x@HC193 Hydrothermal method Na–S ∼3 415.7 mA h g−1@2 A g−1 1257.3 mA h g−1@0.1 A g−1 100 cycles@0.1 A g−1 514 mA h g−1 @0.1 A g−1 0.591 ∼100
CNT/MoS2-Co194 Hydrothermal method and annealing Li–S 1.5 634 mA h g−1@5.0 C 1043 mA h g−1 @1.0 C 800 cycles@1 C 486 mA h g−1 @1 C 0.066 ∼99
HCS/MoS2195 Hydrothermal reaction Na–S 0.5–1 476 mA h g−1@2 C 1090 mA h g−1 @1 C 1000 cycles@1 C 246 mA h g−1@1 C 0.077 ∼100
V-MoS2-CNF196 Sulfurization of MoO3 on CNF Li–S 2 1066 mA h g−1@1 C 1068 mA h g−1 @0.5 C 300 cycles@0.5 C 800 mA h g−1 @0.5 C 0.08 ∼97
MoS2−x/rGO198 Co-dispersion and thermal reduction Li–S ∼0.9 826.5 mA h g−1@8 C 1159.9 mA h g−1@0.5 C 500 cycles@0.5 C 628.2 mA h g−1 @0.5 C 0.083 99.6
LE-MoS2199 Solvothermal reaction and annealing Li–S 2.2 642 mA h g−1@2.0 C 1257 mA h g−1 @1 C 500 cycles@1 C 887 mA h g−1@1 C 0.06 ∼100
1T-phase MoS2 monolayer200 Conversion of Mo1.33CTx in H2S Li–S N/A 532 mA h g−1@5 C ∼1050 mA h g−1@0.5 C 100 cycles@0.5 C 736 mA h g−1 @0.5 C 0.3 ∼100
MXene/1T-2H MoS2-C201 Hydrothermal reaction Li–S 1 677.2 mA h g−1@2 C 1014.1 mA h g−1@0.5 C 300 cycles@0.5 C 799.3 mA h g−1 @0.5 C 0.07 ∼99
FM@G202 One-pot solvothermal reaction Li–S ∼1 823 mA h g−1@5 C 1057 mA h g−1 @1.0 C 500 cycles@1 C 752 mA h g−1 @1 C 0.058 99.6
1T-phase MoS2 nanodots203 Li-intercalation of ground 2H-phase MoS2 Li–S 12.9 883 mA h g−1@4 C 11.3 mA h cm−2@0.05 C 300 cycles@0.05 C 9.4 mA h cm−2 @0.05 C 0.056 N/A
CF@2H/1T MoS2204 One-step hydrothermal method Li–S 1.5 501 mA h g−1@5 C 1048 mA h g−1 @0.5 C 600 cycles@0.5 C 616 mA h g−1 @0.5 C 0.069 ∼99.8
Co–MoS2-G205 Hydrothermal method Li–S 1 941 mA h g−1@2 C 1020 mA h g−1 @1 C 1000 cycles@1 C 723 mA h g−1@1 C 0.029 ∼99.7
P0.1-Mo0.9Ni0.1S2/CNT206 Hydrothermal process and calcination Li–S 1.5 645 mA h g−1@2 C 742 mA h g−1 @1 C 500 cycles@1 C 463 mA h g−1@1 C 0.075 97
2D Mo0.5W0.5S2 alloy207 Co-sputtering and sulfurization Li–S 1.5–2 ∼810 mA h g−1@1 C ∼1353 mA h g−1@0.5 C 400 cycles@0.5 C 790 mA h g−1 @0.5 C ∼0.10 >99
D-MoWS-CNF208 Hydrothermal and vapor etching method Li–S 2 903 mA h g−1@2 C ∼1040 mA h g−1@1 C 1000 cycles@1 C ∼510 mA h g−1 @1 C 0.051 99
TiN/1T-MoS2209 Hydrothermal, nitridation, and lithiation methods Li–S 2.1 689 mA h g−1@2 C ∼775 mA h g−1 @1 C 800 cycles@1 C 459 mA h g−1@1 C 0.051 ∼100
MoS2−x-Co9S8-y/rGO210 Hydrothermal and thermal reduction processes Li–S 0.82 710 mA h g−1@3 C ∼900 mA h g−1 @2 C 600 cycles@2 C 580.7 mA h g−1 @2 C 0.06 ∼100
Fe7S8-MoS2@MoS2-NPC211 Phosphomolybdic acid-etching and vulcanization Li–S 2.45 674.6 mA h g−1@4 C 1250.5 mA h g−1@0.2 C 300 cycles@0.2 C 1000.2 mA h g−1 @0.2 C 0.067 ∼99
MoS2–MoN212 Hydrothermal process and nitridation Li–S 1.2 674 mA h g−1@6 C 872.8 mA h g−1 @1 C 1000 cycles@1 C 520 mA h g−1@1 C 0.039 ∼100
3DG/TM213 Hydrothermal process Li–S 10 613 mA h g−1@2 C ∼908 mA h g−1 @1 C 500 cycles@1 C ∼538 mA h g−1 @1 C 0.08 ∼99.7
C@MoS2214 Solvothermal method Li–S 1.3 554.2 mA h g−1 @5 C 752.5 mA h g−1 @1 C 1000 cycles@1 C 500 mA h g−1@1 C 0.03 ∼98
C@SnO2/TMS215 Hydrothermal process Li–S 2.75 860 mA h g−1@5 C 710 mA h g−1 @5 C 4000 cycles@5 C 448 mA h g−1@5 C 0.009 ∼99


Moreover, the heterogeneous or hybrid structures with metal (e.g., cobalt) or carbon (e.g., carbon spheres, carbon frameworks, and graphene) improve the conductivity of MoS2 to accelerate electron transport, while the carbon synergistically facilitates the exposure of the catalytic sites, promotes ionic conductivity, and buffers the volume expansion of the electrode during repeated discharging/charging, resulting in higher activity and more stable cycling capability.193–195,213,214 Last but not least, non-two-dimensional molybdenum sulfides may contribute to the superb catalytic properties of hybrid catalysts in metal-sulfide batteries. Wang and coworkers constructed a layered C@SnO2/1T-MoS2 array electrode, covered by MoS2 comprising the abundant amorphous state and a weakly crystalline 1T phase, to serve as the sulfur host for Li–S batteries.215 The MoS2 nanosheet and crystalline SnO2 form a heterostructure, where MoS2 expedites Li+ ion diffusion due to its lower barrier and catalyzes LiPS conversion on its abundant defect sites, while SnO2 enables the smooth Li+ transport and strong adsorption of LiSx and also suppresses the volume expansion of sulfur electrode with large pores. As a result, a battery with the C@SnO2/TMS/S cathode delivered a high specific capacity of 1500 mA h g−1 at 0.2 C and exceptional cycling stability with a capacity decay rate of 0.009% per cycle over 4000 cycles at 5 C. These works will inspire more studies on non-two-dimensional MoS2 catalysts for metal–sulfur batteries.

4.4. Metal–oxygen/air batteries

Beyond the metal–sulfur battery applications, MoS2 can also be used as catalysts for oxygen or air cathodes in metal–O2 or metal–air batteries. Li–O2 batteries (LOBs) display the highest energy density of 3500 W h kg−1 based on the ideal electrochemical reaction of 2Li + O2 → Li2O2,216 much higher than those of conventional lithium-ion batteries (100–265 W h kg−1)217 and thus LOBs are considered as one of the most promising next-generation energy storage devices. Zn–air batteries (ZABs) are also potential alternative to LIBs because of their high theoretical energy density (1086 W h kg−1), very low cost, and a high degree of safety.218 Nevertheless, these metal–O2/air batteries consisting of a metal anode and an oxygen or air cathode face many challenges such as sluggish conversion kinetics between oxygen and oxides (e.g., O22−, OH) via OER and ORR reactions. Upgrading MoS2 catalysts using engineering approaches such as heteroatom doping, hybridization, phase modulation, alloying and vacancy defects is effective for advanced metal–O2/air batteries.219–222 In the next section, we will first discuss the challenges in LOBs and solutions with the aid of defective MoS2 catalysts, where defects related to edge sites, vacancies, phase modulation, amorphous states, hybridization, doping, and alloying structure accelerate the electrode reaction in Li–O2 batteries and then we will discuss the difficulties of ZABs and strategies by means of defective MoS2-based heterostructure catalysts.

LOBs face multiple challenges, including large polarization, insufficient capacity, poor round-trip efficiency, and poor cyclability, mainly due to sluggish reactions, parasitic reactions, and poor conductivity.216,219,220,223–225 One of the key approaches for boosting battery performance is to enhance the ORR/OER with catalysts during the discharge/charge process (2Li+ + O2 + 2e → Li2O2, E0 = 2.96 V), thus alleviating the slow kinetics of Li2O2 conversion and suppressing side reactions. MoS2 is considered one of the promising cathode catalysts for OER/ORR.37,225 Compared to MoS2 powder, two-dimensional MoS2 nanoflakes exhibit better catalytic activity during the discharging and charging of LOBs, with the edge sites responsible for catalytic reactions, while the basal plane favors the deposition of discharge products.225 Asadi and coworkers fabricated MoS2 nanoflakes terminated with Mo edges, which had a high density of electrons and exhibited superior catalytic activity for the ORR and the OER.37 The material also showed high round-trip efficiency (∼80%) and up to 50 cycle reversibility when used as a cathode in a Swagelok cell with an ionic liquid (IL, BMIM-BF4) electrolyte (Fig. 19a). The elevated performance is attributed to the exposed Mo edge active sites binding to O2, as well as the IL electrolyte partially covering the Mo edge and facilitating the dissociation of O2 (Fig. 19b) followed by the formation of Li2O2via disproportionation of Li2O (i.e., 2LiO2 → Li2O2 + O2). The authors further disclosed that O2 has stronger interactions than Li2O2 on the Mo edge, while the Li2O2 monomer shows a rather strong interaction with the basal plane by DFT calculations, suggesting the preferable nucleation and growth of Li2O2 on the basal plane (Fig. 19c).226 With the MoS2 nanoflake cathode, ionic liquid/DMSO electrolyte, and Li2CO3-based protective anode, the Swagelok battery experienced a cycle life >500 cycles in an air-like atmosphere (Fig. 19d).


image file: d3mh00462g-f19.tif
Fig. 19 (a) Charging and discharging voltage profiles of the Li–O2 Swagelok battery using MoS2 nanoflakes and an ionic liquid electrolyte saturated with LiTFSI. The inset shows the discharge capacity retention over 50 cycles versus the number of cycles. (b) Illustration of the oxygen reduction reaction on MoS2 nanoflakes in ionic liquid ([EMIM][BF4]), where EMIM+ molecules bind strongly to the negatively charged Mo edge (state 1) to form an EMIM+-covered Mo edge, exposing single-atom Mo sites to the solvent (state 2). O2 then binds to the single-atom Mo sites to form O2 by charge transfer (stage 3). The yellow, green and red balls represent S, Mo and O atoms. The structural model of EMIM+ is shown by gray (for C), blue (for N) and white (for H) balls. (a and b) Reproduced with permission.37 Copyright 2016, American Chemical Society. (c) Calculated conformations of Li2O2 binding on MoS2 nanoribbon. Li2O2 monomer shows a pretty strong binding interaction of −1.02 (left), −1.08 (middle) and −0.83 eV (right) on the basal plane. (d) Discharge–charge voltage profiles for the lithium–air battery system with a lithium carbonate-based protected anode, a MoS2 cathode, and an ionic liquid/ dimethyl sulfoxide electrolyte. The inset shows the capacity versus the number of cycles. (c and d) Reproduced with permission.226 Copyright 2018, Springer Nature. (e) Schematic of the charging and discharging processes of GF-CNT@MoS2 (i.e., amorphous MoS2 deposited on CNT grown on a 3D graphite foam GF) integrated cathode in a Li–O2 battery. (f) Galvanostatic discharge/charge profiles of a Swagelok-type cell with Li metal anode, free-standing GF-CNT@MoS2 cathode, and a glass fiber separator impregnated with the electrolyte at different current densities with a cut-off capacity of 500 mA h g−1. (e and f) Reproduced with permission.216 Copyright 2019, John Wiley & Sons Inc.

Hybridization of MoS2 nanosheets with carbon materials (e.g., CNTs, hollow carbon spheres, and graphene aerogel) is an attractive strategy because of the better electrical conductivity, higher mass transport capability, and expanded storage space to accommodate reaction products.223,224,227,228 In the hybrid, MoS2 increases OER and ORR activity, and suppresses side reactions between the carbonaceous material and reactive oxygen species or electrolyte.216,224,228 Liu and coworkers fabricated sulfur-deficient MoS2−x nanoflakes on microporous carbon, which delivered a smaller polarization gap (0.59 V), a higher discharge capacity (8851 mA h g−1) at a current density of 500 mA g−1, and better reversibility than MoS2 (0.87 V; 6312mA h g−1) in Li–O2 batteries.220 The enhanced performance is attributable to the introduced S-vacancy that offers abundant active centers and the hierarchical carbon structure that allows free diffusion of ions and gases. Sadighi and coworkers developed a 1T-MoS2/CNT structure by hybridizing 1T-phase MoS2 obtained from a Na+ intercalation and exfoliation method with functionalized carbon nanotubes.219 Thanks to the enhanced adsorption energy of O2 and Li2O2 on the catalytically active basal plane of 1T-phase MoS2 as revealed by DFT results, as well as the enhanced electrical conductivity and catalytic activity for ORR and OER, the Li–O2 batteries with 1T-MoS2/CNT as the cathode, exhibited remarkable performance with a high reversible capacity of 500 mA h g−1 for more than 100 cycles at a current density of 200 mA g−1. Li and colleagues fabricated core–shell MoS2−x@CNTs composite, where the charge redistribution on MoS2 nanosheets after the introduction of S-vacancy increases the number of active sites, while the CNT network facilitates the mass transfer and relieves the volume changes during cycling, thus also enabling improved battery performance.229 Song and coworkers reported a thin amorphous MoS2 film (≈5 nm) deposited on carbon nanotube forest (GF-CNT/MoS2) by atomic layer deposition for LOB.216 The disordered MoS2 provides abundant active edge sites for OER/ORR and suppresses parasitic reactions, while the conductive carbon scaffold provides sufficient space for Li2O2 and easy access to the electrolyte (Fig. 19e). As a result, these synergistic effects enabled the GF-CNT/MoS2 electrode to achieve a small overpotential gap (0.58 V) (Fig. 19f) and a high energy efficiency (83%) at 250 mA g−1, a high capacity (4844 mA h g−1) at 500 mA g−1, and a long cycling to 190 cycles.

The hybridization of transition-metal compounds with MoS2 acts on the OER/ORR by affecting the electronic structure, adsorption energy, and carrier transfer. Li and colleagues synthesized decorated MoS2/CoS2 nanorod heterostructures and the coupling between MoS2 with favorable catalytic activity and CoS2 with high electrical conductivity accelerated the electron transfer, exposed numerous edge active sites, and increased catalytic kinetics in the ORR/OER cycling.230 In addition, MoS2 and CoS2 show a high level of lattice matching with Li2O2, which promotes the heteroepitaxial growth of Li2O2. Accordingly, the MoS2/CoS2 heterostructure cathode exhibits high ORR and OER current densities, superior discharge capacity (10495 mA h g−1), along with low overpotentials in discharge–charge curves. Wen and colleagues fabricated MoS2/NiS2 heterostructures with low-coordination atoms (LCAs) derived from S-vacancy at the hetero-interface.231 The d-band center of MoS2/NiS2 (−1.067 eV) is closer to the Fermi level compared to MoS2 (−1.839 eV) and NiS2 (−2.437 eV), which is attributed to the strong electronic interaction between LCAs and the adjacent atoms, leading to enhanced adsorption strength to oxygen intermediates and thus elevated catalytic activity for oxygen redox reactions. Combining with other merits, such as improved charge transfer at the interface, the MoS2/NiS2-based Li–O2 batteries exhibited a high discharge capacity of 12[thin space (1/6-em)]377.4 mA h g−1 (at 200 mA g−1) and a cycling lifetime more than 1000 h in the experiment, revealing the great potential of the heterostructure engineering approach for the design of novel electrocatalysts.

Various other engineering strategies, including doping, alloying, and vacancy engineering, have also been proven to be effective. Doped MoS2 also accelerates the catalytic process and elevates Li–O2 battery's performance in the absence of carbon hybrids. Cao and coworkers synthesized nitrogen doped N–MoS2 with S-vacancies by calcining MoS2 with a nitrogen source.221 As measured by EPR and XPS, the N-to-S substitution resulted in the deficiency of S and the formation of a Mo–N bond, providing more conducive adsorption sites to Li+, more exposed active centers, and promoted reaction kinetics of the ORR and OER. As a result, Li–O2 batteries using the N–MoS2 electrocatalyst delivered a higher discharge-specific capacity (5205.5 mA h g−1) than that with pristine MoS2 (2600.3 mA h g−1) at a current density of 100 mA g−1, as well as cycling stability of 224 cycles at a current density of 200 mA g−1. Zhang and coworkers used a metastable MoSSe solid solution to accelerate the OER/ORR process.222 Poor control of the interlayer vdW forces and different bond lengths of Mo–S and Mo–Se led to high-degree out-of-plane and in-plane lattice distortions in the metastable MoSSe as observed in the HAADF-STEM images (Fig. 20a), from which the calculated strain maps further illustrate the wave-like lattice expansion along the y-direction and slight in-plane lattice distortion along the x-direction of the (002) plane. In addition, the chemical environment is also modulated as shown by XPS, where a decrease in the binding energy of S 2p and an increase in the binding energy of Se 2p were observed. Consequently, Li–O2 batteries with the MoSSe electrode demonstrated a high discharge capacity of 708 mA h g−1 with a small overpotential gap (0.66 V, Fig. 20b) at a current density of 50 mA g−1, as well as reversibility for 30 cycles. Sun and coworkers reported rich-edge S-vacancy MoS2 quantum dots (MoS2 QDs) prepared by Li+-clipping of MoS2 nanosheets using a top-down strategy.232 Compared to the bulk MoS2 electrode used for Li–O2 batteries, the QD electrode (∼3 nm) had a higher discharge capacity (11[thin space (1/6-em)]786 mA h g−1) (Fig. 20c), a smaller overpotential gap (1.04 V), and better cycling stability (230 cycles). DFT calculations revealed that the enhanced performance could be attributed to the edge with S-vacancies, which favored the adsorption and formation of LiO2 and Li2O2 due to the strong adsorption energy and thus accelerated the catalytic ORR and OER processes due to the reduced overpotential. In addition to the above methods, more engineering approaches deserve further exploration, such as hybridization with metals,233,234 incorporation of reaction intermediates,225,235 and design of novel electrode structures for achieving higher OER/ORR performance.226


image file: d3mh00462g-f20.tif
Fig. 20 (a) HAADF-STEM images of the layers (1) and basal plane (2) of the 3D M-MoSSe. Strain map taken along the xx (3) and yy (4) directions calculated from the HAADF image in (2). (b) First-cycle discharge–charge profiles of MoS2, MoSe2, 3D S-MoSSe (3D stable MoSSe solid solution) and 3D M-MoSSe (3D metastable MoSSe solid solution) cathode. The overpotential of 3D M-MoSSe is indicated. (a and b) Reproduced with permission.222 Copyright 2017, American Chemical Society. (c) The initial discharge profiles of the bulk MoS2 (black) and MoS2 QDs (red) cathodes. The discharge capacity of MoS2 QD electrode is much more than that of bulk MoS2. Reproduced with permission.232 Copyright 2019, Elsevier. (d) ORR and (e) OER free energy diagrams of MoS2, MoS2@Fe–N–C surface and MoS2@Fe–N–C interface at U = 0 V and U = 1.23 V, respectively. (f) Charge–discharge curves of wearable ZAB based on MoS2@Fe–N–C NSs (pink line) or Pt/C + Ir/C (blue line) cathode catalysts under repeated folding and releasing conditions. (d–f) Reproduced with permission.218 Copyright 2021, National Academy of Science.

Zinc–air batteries hold great promise for large-scale energy storage owing to their great cost-effectiveness, excellent intrinsic safety, high-abundance of zinc, and high-energy density (1086 W h kg−1).218 The development of bifunctional electrocatalysts to accelerate the sluggish ORR/OER on the air cathode during the discharge/charge process can reduce electrode polarization and improve energy efficiency, increasing capacity and cycling stability. MoS2-based hetero-interfaces have demonstrated high activity for OER/ORR in ZABs.32,236 Yan and colleagues crafted a MoS2@Fe–N–C heterostructure electrocatalyst with atomically dispersed Fe-N4 coupled to MoS2 at the interface, where MoS2 and N could transfer their electrons to the O2* and OH* intermediates to generate OOH*.218 According to the calculated free energy diagrams, MoS2@Fe–N–C (interface) has the lowest reaction barriers for the ORR and OER than those of MoS2 and MoS2@Fe–N–C (surface), indicating that the interface contributes to the enhanced electrocatalytic activity (Fig. 20d and e). The elevated activity of MoS2@Fe–N–C was also confirmed experimentally by a superior ORR half-wave potential (0.84 V) and OER overpotential (360 mV) at 10 mA cm−2. Furthermore, the wearable ZAB with MoS2@Fe–N–C as the air cathode displayed a high specific capacity of 442 mA h g−1Zn, an improved power density of 78 mW cm−2, an outstanding cycling stability of 50 cycles at a current density of 5 mA cm−2, and great rechargeability even with mechanical bending and releasing during the charge/discharge cycle (Fig. 20f). Amiinu and coworkers used a Mo–N/C@MoS2 electrocatalyst with multifunctional catalytic activity for ZABs, which achieved a high power density (196.4 mW cm−2) and voltaic efficiency (≈ 63% at 5 mA cm−2), as well as excellent cycling stability of up to 48 h at 25 mA cm−2.236 The elevated performance is attributed to the synergy of the exposed edges of MoS2, highly active Mo–N coupling centers at the interface of N/C and MoS2, N-induced activated carbon, improved electrical conductivity at the interface, and porous framework with diffusion channels. In addition, when employed as an air electrode, MoS2 with graphene (e.g., MoS2/graphene heterolayer),237 MoS2 with SnS (e.g., MoS2–SnS heterostructure),238 and MoS2 with Co9S8 (e.g., Co9S8@MoS2 core–shell heterostructure)32 also show enhanced ORR/OER activity (see Table 10 for the comparison of catalytic performance of the defective MoS2 materials), robust cycling stability, high power density, and improved cell voltage, due to the synergistic effect of the composition, interface, and defects. It is also noteworthy that, echoing these success in metal–S and metal–O2/air batteries, defective MoS2 is actively broadening its catalytic applications in fuel cells,239 metal–CO2240,241 and other batteries, which also deserves attention.

Table 10 Summary of metal–O2/air battery performance of the defective MoS2-based catalysts
CatalystRef. Method Metal–O2/air battery ORR performance OER performance Electrolyte Polarization gap Capacity Cycling stability
Co9S8@MoS232 Solvothermal approach and annealing Zn–air 0.884 VRHE of CV peak 10 mA cm−2@η = 143 mVRHE KOH/Zn(Ac)2 N/A N/A 120 cycles@10 mA cm−2
MoS2 NFs37 Liquid exfoliation Li−O2 10.5 mA cm−2@2 VLi/Li+ 5.04 mA cm−2@4.2 VLi/Li+ EMIM-BF4/LiTFSI 0.8 V@0.1 mA cm−2 1250 mA h g−1@0.1 mA cm−2 50 cycles@0.1 mA cm−2
GF-CNT@MoS2216 ALD Li−O2 ∼3.1 V Li/Li+ onset potential ∼3.8 V Li/Li+ onset potential LiTFSI/TEGDME 0.58 V@250 mA g−1 4844 mA h g−1@500 mA g−1 190 cycles@500 mA h g−1
MoS2@Fe–N–C NSs218 Hydrothermal and adsorption Zn–air 0.84 VRHE half-wave potential 10 mA cm−2@η = 360 mVRHE in KOH KOH/Zn(Ac)2 N/A 442 mA h g−1Zn@5 mA cm−2 50 cycles@5 mA cm−2
2D 1T-MoS2/CNT219 Intercalation and exfoliation Li−O2 0.81 VRHE half-wave potential 10 mA cm−2@∼1.52 VRHE LiTFSI/TEGDME ∼1.33 V@250 mA g−1 500 mA h g−1@200 mA g−1 100 cycles@200 mA g−1
MoS2−x /carbon220 Hydrothermal reaction Li−O2 N/A N/A LiTFSI/DMSO/EMIM-BF4 0.59 V@500 mA g−1 8851 mA h g−1@500 mA g−1 123 cycles@500 mA g−1
N-MoS2221 Hydrothermal method and calcination Li−O2 0.81 VRHE half-wave potential 10 mA cm−2@1.56 VRHE LiTFSI/TEGDME 0.82 V@100 mA g−1 5205.5 mA h g−1@100 mA g−1 224 cycles@200 mA g−1
3D M-MoSSe222 Hydrothermal reaction Li−O2 N/A N/A LiClO4/DMSO 0.66 V@50 mA g−1 730 mA h g−1@50 mA g−1 30 cycles@50 mA g−1
MoSx/HRG223 Hydrothermal reaction and freeze-drying Li−O2 N/A N/A LiTFSI/TEGDME 1.5 V@0.05 mA cm−2 6678.4 mA h g−1@0.05 mA cm−2 30 cycles@0.1 mA cm−2
Co-N/C@C-MoS2224 Hydrothermal method Li−O2 0.8 VRHE half-wave potential 10 mA cm−2@1.65 VRHE N/A 1.05 V@100 mA g−1 21197 mA h g−1@100 mA g−1 332 cycles@500 mA g−1
2D MoS2 NFs225 Liquid-phase exfoliation Li−air N/A N/A LiTFSI/InBr3/DMSO/EMIM-BF4 N/A 1250 mA h g−1@1 A g−1 240 cycles @1 A g−1
MoS2 nanoflakes226 Liquid exfoliation Li–air N/A N/A LiTFSI/EMIM-BF4/DMSO 0.88 V@500 mA g−1 500 mA h g−1@500 mA g−1 700 cycles@500 mA g−1
MoS2/CNTs227 Hydrothermal growth Li−O2 ∼2.5 V Li/Li+ onset potential ∼3.9 V Li/Li+ onset potential LiTFSI/TEGDME 1.34 V@200 mA g−1 6904 mA h g−1@200 mA g−1 132 cycles@200 mA g−1
MoS2/HCS228 Hydrothermal reaction Li−O2 ∼2.7 V Li/Li+ onset potential ∼4.2 V Li/Li+ onset potential LiTFSI/TEGDME N/A 4010 mA h g−1@200 mA g−1 104 cycles@200 mA g−1
MoS2−x@CNTs229 Hydrothermal and reduction reactions Li−O2 ∼2.7 V Li/Li+ onset potential ∼3.9 V Li/Li+ onset potential LiTFSI/TEGDME 1.25 V@200 mA g−1 19[thin space (1/6-em)]989 mA h g−1@200 mA g−1 666 cycles@1000 mA g−1
MoS2/CoS2230 Hydrothermal and sulfuration processes Li−O2 ∼2.7 V Li/Li+ onset potential ∼4.2 V Li/Li+ onset potential LiTFSI/TEGDME N/A 10495 mA h g−1@100 mA g−1 120 cycles@200 mA g−1
MoS2/NiS2231 Hydrothermal method Li−O2 2.89 V Li/Li+ onset potential ∼4.0 V Li/Li+ onset potential LiTFSI/TEGDME 0.34 V@100 mA g−1 12377 mA h g−1@200 mA g−1 1101 h@0.1 mV s−1
Rich-edge S-vacancy MoS2 QDs232 Ultrasonic treatment Li−O2 ∼2.7 V Li/Li+ onset potential ∼3.9 V Li/Li+ onset potential LiTFSI/TEGDME 1.04 V@500 mA g−1 11786 mA h g−1@500 mA g−1 230 cycles@500 mA g−1
MoS2/AuNP nanohybrid234 Hydrothermal synthesis Li−O2 ∼2.9 V Li/Li+ onset potential ∼4.1 V Li/Li+ onset potential LiCF3SO3/TEGDME 1.49 V@70 mA g−1 4336 mA h g−1@70 mA g−1 50 cycles@300 mA g−1
rGO-MoS2235 Hydrothermal reaction Hybrid Li−ion/Li–O2 N/A N/A LiTFSI/DME/LiI 0.02 V@50 mA g−1 ∼440 mA h g−1@50 mA g−1 500 cycles@250 mA g−1
Mo–N/C@MoS2236 Hydrothermal method and carbonization Zn–air 0.81 VRHE half-wave potential 10 mA cm−2@η = 0.39 VRHE KOH/Zn(Ac)2 0.75 V@5 mA cm−2 846 W h kg−1Zn@5 mA cm−2 12 cycles@25 mA cm−2
MoS2/Gr237 Stirring Zn–air Strong ORR peak at 0.253 VAg/AgCl N/A KOH N/A 130 W h kg−1 N/A
MoS2-SnS/NPC238 Freeze-drying and calcination Zn–air 0.802 VRHE half-wave potential ∼1.8 VRHE onset potential KOH ∼0.83 V@5 mA cm−2 797 mA h g−1@10 mA cm−2 288 cycles@5 mA cm−2


4.5. Hydrodesulfurization reaction

Catalytic desulfurization for petroleum upgrading via a MoS2 catalyst is an important industrial application that has been commercialized. Sulfur-containing compounds in petroleum, such as mercaptan, thiophene, benzo-thiophene, and dibenzothiophene cause the release of sulfides during combustion. Fuels with a low sulfur content (<10 ppm level in petrochemicals), consequently, are required in many countries to meet the environmental needs,242,243 driving the development of advanced hydrodesulfurization (HDS) materials. It is worth mentioning that H2S generated from petroleum desulfurization is transformed into S or S-based chemicals through industrial recovery and conversion plants, eliminating the concern about by-products from HDS. MoS2 particles and their hybrids, MoS2-based heterostructures, Co(Ni)-promoted MoS2, MoS2 with single-atom anchoring, etc. have emerged as the main catalysts for the HDS processes.242,244–248 Among the diverse defect types, here we focus and briefly discuss the defective MoS2 with S-vacancies because the S-vacancy sites accelerate the adsorption of sulfur-containing molecules and promote the cleavage of C–S bonds in compounds.243 Much effort revolving around the vacancy formation, catalysis process, and structure maintenance of defective S-vacancies has been made to deepen the understanding of the underlying mechanism.243,249,250 In the next section, we will first discuss the role of S-vacancies in the catalytic process, then the formation mechanism of the S-vacancies, and finally their structural stability.

For the desulfurization process of petroleum, S-vacancies provide chemisorption sites with a lowered activation energy barrier of the HDS reaction, and S-vacancies also induce the electron-deficient properties of the surrounding Mo atoms to further reinforce the adsorption of S-containing compounds.249 Moreover, the location of S-vacancy sites affects the catalytic reaction pathways and/or the products. Zheng and coworkers pointed out that the HDS reactions follow either the hydrogenation (HYD) route at the S edge or the direct desulfurization (DDS) route at the Mo edge (Fig. 21a).243 As for the products, the formation of butane bears relatively high reaction barriers, while the production of butene is much more favorable along their respective routes. Salazar and coauthors demonstrated experimentally that the location of S-vacancies (Vs) in the nanoparticle affects the geometric configuration of the generated S-vacancies and also the subsequent adsorption mode of thiophene.242 After the occupation of S-vacancies by thiophene, the atom-resolved scanning tunneling microscopy (STM) images display that thiophene can directly adsorb at the corner (C) site, while its adsorption on adjacent (A) and middle (M) sites is accompanied by S displacement to form the S2 dimer, and thiophene adsorbs at the double VS vacancies to reduce the effects of the steric constraints (Fig. 21b and c). Tuxen and colleagues suggested that the size of MoS2 nanoparticles leads to differences in dibenzothiophene (DBT) adsorption.251 Larger Mo-edge-terminated nanoparticles energetically favor the formation of S-vacancies at the edge, while smaller S-edge-terminated nanoparticles favor the formation of S-vacancies at the corner site. During DBT desulfurization, compared with the weak DBT affinity due to steric hindrance at the edge of large-sized MoS2 nanoparticles, small-sized nanoparticles with S-vacancies at the corner are favorable for DBT adsorption, attributed to the accessible Mo site (Fig. 21d and e).


image file: d3mh00462g-f21.tif
Fig. 21 (a) Schematic of the desulfurization reaction depending on the location of vacancies, where S-vacancies at the S-edge follow the HYD route, while those at the Mo-edge follow the DDS route. Reproduced with permission.243 Copyright 2017, Elsevier. (b) Top-view ball model of MoS2 (left) and side-view ball model of the Mo-edge with a 50% nominal coverage for three edge lengths (right). The corner (C), adjacent to corner (A), and middle (M) positions are marked, respectively. (c) STM images of three cases of active MoS2 nanoparticles after exposure to thiophene at 300 K, shown in superimposed mode, which display the location of thiophene molecules (marked in yellow) and sulfur dimer formation (marked in red), with the direction of atomic displacement indicated by green dashed arrows. The scale bar is 1 nm. (b and c) Reproduced with permission.242 Copyright 2020, Springer Nature. (d) Illustrations of the DBT adsorption configurations at S-vacancies on Mo-edge-terminated (left) and S-edge-terminated (right) MoS2 nanoparticles. The red curve near the Mo-edge S-vacancy shows steric hindrance. (e) STM images of DBT molecules on the Mo-edge or S-edge corner site with the adsorption models superimposed on the STM images. The distances between the S atom of DBT and the corner Mo atom are indicated. (d and e) Reproduced with permission.251 Copyright 2010, American Chemical Society. (f) Ab initio thermodynamics phase diagram of the MoS2 edge structures in H2/H2S mixtures. ΔμS and ΔμH designate the entropic parts of the chemical potentials of S and H atoms in the gas phase, respectively. The experimental conditions during catalyst preparation, during 1 bar hydrogen, and during HDS are indicated in blue. The ball models represent side views of the structures present in the phase diagram. Reproduced with permission.250 Copyright 2019, Springer Nature.

As the main active sites, the S-vacancies in MoS2 would be generated directly under the HDS conditions with a high temperature and hydrogen atmosphere and the formation mechanism of vacancies can be deduced from thermodynamic and kinetic perspectives. Shang and coworkers investigated the generation of sulfur vacancies over the Mo edge (−1010) with 50% S coverage and the S edge (10–10) with 100% S coverage and found the formation of S-vacancies on the S edge proceeds spontaneously with a negative reaction Gibbs free energy (ΔG < 0) while it is almost inhibited at the Mo edge (ΔG > 0).249 The kinetic steps for vacancy formation demonstrate that the H2 molecules prefer homolytic cleavage to form double S–H bonds on two S atoms and the formation of adsorbed H2S from the dissociated H atoms on the S edge is the rate-controlling process on account of its highest energy barrier. In contrast, Paul and colleague suggested that the extraction of S atoms may occur at the Mo edge, where the activation energy of its rate determining step (<1 eV) is lower than that of the S edge (>1.3 eV).252 The heterolytic dissociation of H2 molecules to form one S–H and one Mo–H group is supposed to be the rate-determining step at the Mo edge. The intrinsic endothermic reaction of forming S-vacancies at the Mo edge or S edge suggests an unfavorable thermodynamic process, which requires harsh hydrodesulfurization conditions to facilitate the departure of S atoms in the industrial process. In addition, Zheng and coauthors disclosed that H2 prefers to form S–H and Mo–H species at the Mo edge, whilst to form two S–H species at the S edge after H2 cleavage and the overall process starting with the scission of H2 and ending with the departure of H2S tended to occur at the Mo edge due to its moderate energy barriers.243 Salazar and coworkers calculated VS formation energies of three types of S-vacancies on the corner, adjacent corner, and middle positions, respectively, and found their endothermic processes during vacancy generation.242 In conclusion, the mechanism of S-vacancy formation during hydrodesulfurization is still theoretically controversial.

Finally, in terms of structure maintenance, Mom and colleagues investigated the effect of harsh conditions on the catalyst structure by observing the active MoS2 edge structure under reaction conditions (1 bar, 250 °C).250 Under pure H2 conditions, the particle edge shrinks but maintains the periodic lattice spacing along the edges to accommodate the ensuing CH3SH adsorption, while under hydrodesulfurization conditions, the edge structure changes again since a mixture of adsorbed sulfur and CH3SH occupies the dominant edge structure during the desulfurization of CH3SH. Furthermore, the phase diagram theoretically depicts the most stable edge structure of Au-supported MoS2 as a function of ΔμS and ΔμH, which are directly related to temperature, H2S partial pressure, and H2 partial pressure (Fig. 21f). Thus, the active edge sites adapt their sulfur, hydrogen, and hydrocarbon coverages according to the gas environment, thus adjusting the prevalent structure. Last but not least, it is worth mentioning that, similar to HDS, MoS2 catalysts can also be used in hydrodeoxygenation (HDO),253,254 hydrodenitrogenation (HDN),255 CO2 hydrogenation,256 and so on, and the high catalytic activity undoubtedly relies on the sophisticated design of the defective structure.

5. Conclusions and outlook

As one of the most promising candidates of nonprecious catalysts, molybdenum sulfide has many potential applications. Defects play a vital role in enhancing the catalytic activity of molybdenum sulfide-based materials. This review has summarized ongoing research efforts on defective molybdenum sulfide for electrocatalytic applications, including the use of defective 2D molybdenum disulfide (i.e., 2H, 1T, and 3R phases) and defective non-2D molybdenum sulfide (i.e., amorphous MoSx, MoSx clusters) for the HER, NRR, CO2RR, metal–sulfur battery, and metal–O2/air battery applications. We review theoretical computational work to unravel the correlation between defects and catalytic performances of molybdenum sulfide electrocatalytic systems from both thermodynamic and kinetic perspectives and to guide the use of defects to tune materials for achieving enhanced adsorption behavior and optimized catalytic performance. We summarize a range of defect types, including vacancies, strained structures, dislocations, distortions, modulated phases, amorphous states, clusters, and exposed edges, as well as various defect generation methods such as plasma treatment, hydrothermal synthesis, heated growth, heteroatom doping, ion intercalation, hybridization, electrochemical deposition, and solid solution routes. Despite these advances, a few challenges remain:

(1) Lack of methods for precise creation of defects

Although much progress has been made in developing defect generation methods, the quest for more precise fabrication techniques, such as precise control of defect types and sites at the atomic level, is still ongoing. Meanwhile, the coupling of defects has been shown to achieve synergistic enhancement effects. Sophisticated designs are still needed to precisely couple different types of defects to achieve breakthroughs in catalytic performance and thus improve energy conversion efficiency. Factors such as structural stability, electrical conductivity, degree of defects, and controllability of the defect preparation method also need to be considered.

(2) Insufficient depth of mechanism study

Theoretical calculations provide an excellent tool for establishing material–structure correlations and therefore a means for high-throughput screening of materials. For the mechanistic analysis of defects, further optimization of computational methods is needed to achieve more accurate predictions and quantification. Meanwhile, the defect structure is often accompanied by the coupling of various defect types, so the study of a single defect after decoupling, as well as a comprehensive understanding of the catalytic effect of coupled defects, is crucial. Moreover, in situ experimental observations are often used for mechanism verification, such as in situ synchrotron radiation characterization of catalytic active sites, and more in situ characterization studies are needed to increase the accuracy and depth of defect mechanism studies.

(3) Limited demonstration of applications

The catalytic applications of defective molybdenum sulfides are not limited to the HER, NRR, CO2RR, metal–sulfur batteries, and metal–O2/air batteries; there are many more catalytic applications that can be explored. Mature electrochemical applications are being aggressively pursued in practice, and a broader range of electrochemical applications are emerging at the intersection of various disciplines. The catalytic reduction of NO to ammonia and the conversion of organic molecules via MoSx catalyst as mentioned above have not yet been systematically studied. In particular, there is a wide scope for innovation in organic electrocatalysis, so catalytic materials based on defective molybdenum disulfide need to be further explored in these emerging fields of electrocatalysis.

(4) Lack of industrial practice

Molybdenum disulfide catalysts for electrochemical hydrogen production have made considerable progress in fundamental research but the development to date is still very far from industrial applications. For practical applications, the following issues need to be addressed, including further improvement of the activity and stability of the catalytic materials; more efforts in simplifying the defect preparation method, controlling costs of materials and preparation processes, and scaling up of the materials; development of advanced membrane electrode assembly techniques for new catalysts applicable to PEM electrolyzers; and more operational parameters of PEM electrolyzers based on molybdenum disulfide materials, which comprise energy efficiency, gas purity, operating voltage, current density, stability, etc. For example, commercial PEM electrolyzers have achieved operational stability for more than 20[thin space (1/6-em)]000 hours, but current stability based on molybdenum sulfide materials is far from this standard. In addition, other potential industrial applications of electrochemistry need to be actively promoted, such as developing electrochemical reduction of carbon dioxide to produce carbon monoxide or liquid products for carbon capture or storage and improving the catalytic properties of electrode materials for metal–sulfur and metal–oxygen batteries to facilitate their commercialization. Moreover, it is also of great significance to integrate sustainable energy sources such as solar, wind, and tidal power and electrocatalytic applications to achieve electrocatalytic conversion and energy storage of renewable green energy.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This work was supported by Nanyang Technological University under NAP award (M408050000) and Guangdong Basic and Applied Basic Research Foundation (2022A1515110532).

References

  1. B. Han and Y. H. Hu, Energy Sci. Eng., 2016, 4, 285–304 CrossRef CAS .
  2. Y. Fang, J. Pan, J. He, R. Luo, D. Wang, X. Che, K. Bu, W. Zhao, P. Liu, G. Mu, H. Zhang, T. Lin and F. Huang, Angew. Chem., Int. Ed., 2018, 57, 1232–1235 CrossRef CAS PubMed .
  3. A. N. Enyashin, L. Yadgarov, L. Houben, I. Popov, M. Weidenbach, R. Tenne, M. Bar-Sadan and G. Seifert, J. Phys. Chem. C, 2011, 115, 24586–24591 CrossRef CAS .
  4. U. Maitra, U. Gupta, M. De, R. Datta, A. Govindaraj and C. Rao, Angew. Chem., Int. Ed., 2013, 52, 13057–13061 CrossRef CAS PubMed .
  5. S.-Z. Yang, Y. Gong, P. Manchanda, Y.-Y. Zhang, G. Ye, S. Chen, L. Song, S. T. Pantelides, P. M. Ajayan, M. F. Chisholm and W. Zhou, Adv. Mater., 2018, 30, 1803477 CrossRef PubMed .
  6. H. Tan, W. Hu, C. Wang, C. Ma, H. Duan, W. Yan, L. Cai, P. Guo, Z. Sun, Q. Liu, X. Zheng, F. Hu and S. Wei, Small, 2017, 13, 1701389 CrossRef PubMed .
  7. J. Peng, Y. Liu, X. Luo, J. Wu, Y. Lin, Y. Guo, J. Zhao, X. Wu, C. Wu and Y. Xie, Adv. Mater., 2019, 31, 1900568 CrossRef PubMed .
  8. W. Zhao, J. Pan, Y. Fang, X. Che, D. Wang, K. Bu and F. Huang, Chem. – Eur. J., 2018, 24, 15942–15954 CrossRef CAS PubMed .
  9. G. Seifert, H. Terrones, M. Terrones, G. Jungnickel and T. Frauenheim, Phys. Rev. Lett., 2000, 85, 146 CrossRef CAS PubMed .
  10. T. Böker, R. Severin, A. Müller, C. Janowitz, R. Manzke, D. Voß, P. Krüger, A. Mazur and J. Pollmann, Phys. Rev. B: Condens. Matter Mater. Phys., 2001, 64, 235305 CrossRef .
  11. K. F. Mak, C. Lee, J. Hone, J. Shan and T. F. Heinz, Phys. Rev. Lett., 2010, 105, 136805 CrossRef PubMed .
  12. L. Sun, J. Yan, D. Zhan, L. Liu, H. Hu, H. Li, B. K. Tay, J.-L. Kuo, C.-C. Huang, D. W. Hewak, P. S. Lee and Z. X. Shen, Phys. Rev. Lett., 2013, 111, 126801 CrossRef PubMed .
  13. T. Li and G. Galli, J. Phys. Chem. C, 2007, 111, 16192–16196 CrossRef CAS .
  14. O. V. Yazyev and A. Kis, Mater. Today, 2015, 18, 20–30 CrossRef CAS .
  15. A. Splendiani, L. Sun, Y. Zhang, T. Li, J. Kim, C.-Y. Chim, G. Galli and F. Wang, Nano Lett., 2010, 10, 1271–1275 CrossRef CAS PubMed .
  16. Z. Yin, H. Li, H. Li, L. Jiang, Y. Shi, Y. Sun, G. Lu, Q. Zhang, X. Chen and H. Zhang, ACS Nano, 2012, 6, 74–80 CrossRef CAS PubMed .
  17. H. Li, Z. Yin, Q. He, H. Li, X. Huang, G. Lu, D. W. H. Fam, A. I. Y. Tok, Q. Zhang and H. Zhang, Small, 2012, 8, 63–67 CrossRef CAS PubMed .
  18. A. M. Van Der Zande, P. Y. Huang, D. A. Chenet, T. C. Berkelbach, Y. You, G.-H. Lee, T. F. Heinz, D. R. Reichman, D. A. Muller and J. C. Hone, Nat. Mater., 2013, 12, 554–561 CrossRef CAS PubMed .
  19. H. Li, C. Tsai, A. L. Koh, L. Cai, A. W. Contryman, A. H. Fragapane, J. Zhao, H. S. Han, H. C. Manoharan, F. Abild-Pedersen, J. K. Nørskov and X. Zheng, Nat. Mater., 2016, 15, 48–53 CrossRef CAS PubMed .
  20. H. Li, A. W. Contryman, X. Qian, S. M. Ardakani, Y. Gong, X. Wang, J. M. Weisse, C. H. Lee, J. Zhao, P. M. Ajayan, J. Li, H. C. Manoharan and X. Zheng, Nat. Commun., 2015, 6, 7381 CrossRef CAS PubMed .
  21. B. Mohanty, M. Ghorbani-Asl, S. Kretschmer, A. Ghosh, P. Guha, S. K. Panda, B. Jena, A. V. Krasheninnikov and B. K. Jena, ACS Catal., 2018, 8, 1683–1689 CrossRef CAS .
  22. C. Tang, L. Zhong, B. Zhang, H. F. Wang and Q. Zhang, Adv. Mater., 2018, 30, 1705110 CrossRef PubMed .
  23. M. Asadi, B. Kumar, A. Behranginia, B. A. Rosen, A. Baskin, N. Repnin, D. Pisasale, P. Phillips, W. Zhu, R. Haasch, R. F. Klie, P. Král, J. Abiade and A. Salehi-Khojin, Nat. Commun., 2014, 5, 4470 CrossRef CAS PubMed .
  24. L. Zhang, X. Ji, X. Ren, Y. Ma, X. Shi, Z. Tian, A. M. Asiri, L. Chen, B. Tang and X. Sun, Adv. Mater., 2018, 30, 1800191 CrossRef PubMed .
  25. L. Zhang, J. Liang, Y. Wang, T. Mou, Y. Lin, L. Yue, T. Li, Q. Liu, Y. Luo, N. Li, B. Tang, Y. Liu, S. Gao, A. A. Alshehri, X. Guo, D. Ma and X. Sun, Angew. Chem., Int. Ed., 2021, 60, 25263–25268 CrossRef CAS PubMed .
  26. L. Maachou, K. Qi, E. Petit, Z. Qin, Y. Zhang, D. Cot, V. Flaud, C. Reibel, H. El-Maghrbi, L. Li, P. Miele, D. Kaplan, M. Chhowalla, N. Onofrio and D. Voiry, J. Mater. Chem. A, 2020, 8, 25053–25060 RSC .
  27. D. He, H. Ooka, Y. Kim, Y. Li, F. Jin, S. H. Kim and R. Nakamura, Proc. Natl. Acad. Sci. U. S. A., 2020, 117, 31631–31638 CrossRef CAS PubMed .
  28. Y. Zang, S. Niu, Y. Wu, X. Zheng, J. Cai, J. Ye, Y. Xie, Y. Liu, J. Zhou, J. Zhu, X. Liu, G. Wang and Y. Qian, Nat. Commun., 2019, 10, 1217 CrossRef PubMed .
  29. K. Jiang, M. Luo, Z. Liu, M. Peng, D. Chen, Y.-R. Lu, T.-S. Chan, F. M. de Groot and Y. Tan, Nat. Commun., 2021, 12, 1687 CrossRef CAS PubMed .
  30. X. Huang, H. Xu, D. Cao and D. Cheng, Nano Energy, 2020, 78, 105253 CrossRef CAS .
  31. T. Zhang, Y. Liu, J. Yu, Q. Ye, L. Yang, Y. Li and H. J. Fan, Adv. Mater., 2022, 34, 2202195 CrossRef CAS PubMed .
  32. J. Bai, T. Meng, D. Guo, S. Wang, B. Mao and M. Cao, ACS Appl. Mater. Interfaces, 2018, 10, 1678–1689 CrossRef CAS PubMed .
  33. B. H. R. Suryanto, D. Wang, L. M. Azofra, M. Harb, L. Cavallo, R. Jalili, D. R. G. Mitchell, M. Chatti and D. R. MacFarlane, ACS Energy Lett., 2018, 4, 430–435 CrossRef .
  34. Y. Liu, M. Han, Q. Xiong, S. Zhang, C. Zhao, W. Gong, G. Wang, H. Zhang and H. Zhao, Adv. Energy Mater., 2019, 9, 1803935 CrossRef .
  35. R. Fang, S. Zhao, Z. Sun, D.-W. Wang, H.-M. Cheng and F. Li, Adv. Mater., 2017, 29, 1606823 CrossRef PubMed .
  36. Y. Wang, Y. Lai, J. Chu, Z. Yan, Y. X. Wang, S. L. Chou, H. K. Liu, S. X. Dou, X. Ai, H. Yang and Y. Cao, Adv. Mater., 2021, 33, 2100229 CrossRef CAS PubMed .
  37. M. Asadi, B. Kumar, C. Liu, P. Phillips, P. Yasaei, A. Behranginia, P. Zapol, R. F. Klie, L. A. Curtiss and A. Salehi-Khojin, ACS Nano, 2016, 10, 2167–2175 CrossRef CAS PubMed .
  38. S. E. Skrabalak and K. S. Suslick, J. Am. Chem. Soc., 2005, 127, 9990–9991 CrossRef CAS PubMed .
  39. Y. Yin, J. Han, Y. Zhang, X. Zhang, P. Xu, Q. Yuan, L. Samad, X. Wang, Y. Wang, Z. Zhang, P. Zhang, X. Cao, B. Song and S. Jin, J. Am. Chem. Soc., 2016, 138, 7965–7972 CrossRef CAS PubMed .
  40. W. Wu, C. Niu, C. Wei, Y. Jia, C. Li and Q. Xu, Angew. Chem., Int. Ed., 2019, 58, 2029–2033 CrossRef CAS PubMed .
  41. B. Shang, X. Cui, L. Jiao, K. Qi, Y. Wang, J. Fan, Y. Yue, H. Wang, Q. Bao, X. Fan, S. Wei, W. Song, Z. Cheng, S. Guo and W. Zheng, Nano Lett., 2019, 19, 2758–2764 CrossRef CAS PubMed .
  42. L. Lei, Y. Lun, F. Cao, L. Meng, S. Xing, J. Guo, H. Dong, S. Gu, K. Xu, S. Hussain, Y. Li, Y. Sugawara, F. Pang, W. Ji, J. Hong, R. Xu and Z. Cheng, Nanotechnology, 2021, 32, 465703 CrossRef CAS PubMed .
  43. Y. Zhang, T. Yang, J. Li, Q. Zhang, B. Li and M. Gao, Adv. Funct. Mater., 2023, 33, 2210939 CrossRef CAS .
  44. T. Sun, J. Wang, X. Chi, Y. Lin, Z. Chen, X. Ling, C. Qiu, Y. Xu, L. Song, W. Chen and C. Su, ACS Catal., 2018, 8, 7585–7592 CrossRef CAS .
  45. Q. Ke, X. Zhang, W. Zang, A. M. Elshahawy, Y. Hu, Q. He, S. J. Pennycook, Y. Cai and J. Wang, Small, 2019, 15, 1900131 CrossRef PubMed .
  46. T. Ren, W. Tian, Q. Shen, Z. Yuan, D. Chen, N. Li and J. Lu, Nano Energy, 2021, 90, 106527 CrossRef CAS .
  47. S. Deng, M. Luo, C. Ai, Y. Zhang, B. Liu, L. Huang, Z. Jiang, Q. Zhang, L. Gu, S. Lin, X. Wang, L. Yu, J. Wen, J. Wang, G. Pan, X. Xia and J. Tu, Angew. Chem., Int. Ed., 2019, 58, 16289–16296 CrossRef CAS PubMed .
  48. N. Feng, R. Meng, L. Zu, Y. Feng, C. Peng, J. Huang, G. Liu, B. Chen and J. Yang, Nat. Commun., 2019, 10, 1372 CrossRef PubMed .
  49. Z. Li, H. Li, Z. Yang, X. Lu, S. Ji, M. Zhang, J. H. Horton, H. Ding, Q. Xu, J. Zhu and J. Yu, Small, 2022, 18, 2201092 CrossRef CAS PubMed .
  50. Z. Li, X. Yan, D. He, W. Hu, S. Younan, Z. Ke, M. Patrick, X. Xiao, J. Huang, H. Wu, X. Pan and J. Gu, ACS Catal., 2022, 12, 7687–7695 CrossRef CAS .
  51. X. Wang, L. Pan, J. Yang, B. Li, Y. Y. Liu and Z. Wei, Adv. Mater., 2021, 33, 2006908 CrossRef CAS PubMed .
  52. Y. Zhang, H. Tao, S. Du and X. Yang, ACS Appl. Mater. Interfaces, 2019, 11, 11327–11337 CrossRef CAS PubMed .
  53. D. Escalera-López, Z. Lou and N. V. Rees, Adv. Energy Mater., 2019, 9, 1802614 CrossRef .
  54. D. Escalera-López, Y. Niu, J. Yin, K. Cooke, N. V. Rees and R. E. Palmer, ACS Catal., 2016, 6, 6008–6017 CrossRef PubMed .
  55. J. Kibsgaard, T. F. Jaramillo and F. Besenbacher, Nat. Chem., 2014, 6, 248–253 CrossRef CAS PubMed .
  56. R. Kothari, D. Buddhi and R. Sawhney, Renewable Sustainable Energy Rev., 2008, 12, 553–563 CrossRef CAS .
  57. M. Carmo, D. L. Fritz, J. Mergel and D. Stolten, Int. J. Hydrogen Energy, 2013, 38, 4901–4934 CrossRef CAS .
  58. K. E. Ayers, J. N. Renner, N. Danilovic, J. X. Wang, Y. Zhang, R. Maric and H. Yu, Catal. Today, 2016, 262, 121–132 CrossRef CAS .
  59. B. Hinnemann, P. G. Moses, J. Bonde, K. P. Jørgensen, J. H. Nielsen, S. Horch, I. Chorkendorff and J. K. Nørskov, J. Am. Chem. Soc., 2005, 127, 5308–5309 CrossRef CAS PubMed .
  60. T. F. Jaramillo, K. P. Jørgensen, J. Bonde, J. H. Nielsen, S. Horch and I. Chorkendorff, Science, 2007, 317, 100–102 CrossRef CAS PubMed .
  61. M. A. Lukowski, A. S. Daniel, F. Meng, A. Forticaux, L. Li and S. Jin, J. Am. Chem. Soc., 2013, 135, 10274–10277 CrossRef CAS PubMed .
  62. D. Voiry, R. Fullon, J. Yang, C. de Carvalho Castro e Silva, R. Kappera, I. Bozkurt, D. Kaplan, M. J. Lagos, P. E. Batson, G. Gupta, A. D. Mohite, L. Dong, D. Er, V. B. Shenoy, T. Asefa and M. Chhowalla, Nat. Mater., 2016, 15, 1003–1009 CrossRef CAS PubMed .
  63. Y. Li, H. Wang, L. Xie, Y. Liang, G. Hong and H. Dai, J. Am. Chem. Soc., 2011, 133, 7296–7299 CrossRef CAS PubMed .
  64. G. Li, D. Zhang, Q. Qiao, Y. Yu, D. Peterson, A. Zafar, R. Kumar, S. Curtarolo, F. Hunte, S. Shannon, Y. Zhu, W. Yang and L. Cao, J. Am. Chem. Soc., 2016, 138, 16632–16638 CrossRef CAS PubMed .
  65. C. Tsai, H. Li, S. Park, J. Park, H. S. Han, J. K. Nørskov, X. Zheng and F. Abild-Pedersen, Nat. Commun., 2017, 8, 15113 CrossRef PubMed .
  66. L. Li, Z. 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 PubMed .
  67. H. Li, M. Du, M. J. Mleczko, A. L. Koh, Y. Nishi, E. Pop, A. J. Bard and X. Zheng, J. Am. Chem. Soc., 2016, 138, 5123–5129 CrossRef CAS PubMed .
  68. C.-C. Cheng, A.-Y. Lu, C.-C. Tseng, X. Yang, M. N. Hedhili, M.-C. Chen, K.-H. Wei and L.-J. Li, Nano Energy, 2016, 30, 846–852 CrossRef CAS .
  69. C. Meng, M.-C. Lin, X.-W. Du and Y. Zhou, ACS Sustainable Chem. Eng., 2019, 7, 6999–7003 CrossRef CAS .
  70. G. Ye, Y. Gong, J. Lin, B. Li, Y. He, S. T. Pantelides, W. Zhou, R. Vajtai and P. M. Ajayan, Nano Lett., 2016, 16, 1097–1103 CrossRef CAS PubMed .
  71. J. Yang, Y. Wang, M. J. Lagos, V. Manichev, R. Fullon, X. Song, D. Voiry, S. Chakraborty, W. Zhang, P. E. Batson, L. Feldman, T. Gustafsson and M. Chhowalla, ACS Nano, 2019, 13, 9958–9964 CrossRef CAS PubMed .
  72. C. Wang, H. Lu, K. Tang, Z. Mao, Q. Li, X. Wang and C. Yan, Electrochim. Acta, 2020, 336, 135740 CrossRef CAS .
  73. S. Geng, W. Yang, Y. Liu and Y. Yu, J. Catal., 2020, 391, 91–97 CrossRef CAS .
  74. X. Wang, Y. Zhang, H. Si, Q. Zhang, J. Wu, L. Gao, X. Wei, Y. Sun, Q. Liao, Z. Zhang, K. Ammarah, L. Gu, Z. Kang and Y. Zhang, J. Am. Chem. Soc., 2020, 142, 4298–4308 CrossRef CAS PubMed .
  75. P. Zhang, H. Xiang, L. Tao, H. Dong, Y. Zhou, T. S. Hu, X. Chen, S. Liu, S. Wang and S. Garaj, Nano Energy, 2019, 57, 535–541 CrossRef CAS .
  76. J. Wang, W. Fang, Y. Hu, Y. Zhang, J. Dang, Y. Wu, B. Chen, H. Zhao and Z. Li, Appl. Catal., B, 2021, 298, 120490 CrossRef CAS .
  77. Z. Luo, Y. Ouyang, H. Zhang, M. Xiao, J. Ge, Z. Jiang, J. Wang, D. Tang, X. Cao, C. Liu and W. Xing, Nat. Commun., 2018, 9, 2120 CrossRef PubMed .
  78. J. Xu, G. Shao, X. Tang, F. Lv, H. Xiang, C. Jing, S. Liu, S. Dai, Y. Li, J. Luo and Z. Zhou, Nat. Commun., 2022, 13, 2193 CrossRef CAS PubMed .
  79. S. Park, J. Park, H. Abroshan, L. Zhang, J. K. Kim, J. Zhang, J. Guo, S. Siahrostami and X. Zheng, ACS Energy Lett., 2018, 3, 2685–2693 CrossRef CAS .
  80. Y. Zhao, M. T. Tang, S. Wu, J. Geng, Z. Han, K. Chan, P. Gao and H. Li, J. Catal., 2020, 382, 320–328 CrossRef CAS .
  81. Y. Zhao, J. Hwang, M. T. Tang, H. Chun, X. Wang, H. Zhao, K. Chan, B. Han, P. Gao and H. Li, J. Power Sources, 2020, 456, 227998 CrossRef CAS .
  82. Y. Kang, Surf. Sci., 2021, 704, 121759 CrossRef CAS .
  83. Y. Zhang, Y. Kuwahara, K. Mori and H. Yamashita, Chem. – Eur. J., 2019, 14, 278–285 CAS .
  84. J. Pető, T. Ollár, P. Vancsó, Z. I. Popov, G. Z. Magda, G. Dobrik, C. Hwang, P. B. Sorokin and L. Tapasztó, Nat. Chem., 2018, 10, 1246–1251 CrossRef PubMed .
  85. Y. Zhao, W. Xin, B. Liu, H. Li, Y. Xu and Z. Zhang, Inorg. Chem. Front., 2022, 9, 3461–3469 RSC .
  86. A. M. Z. Tan, C. Freysoldt and R. G. Hennig, Phys. Rev. Mater., 2020, 4, 064004 CrossRef CAS .
  87. J. Ekspong, R. Sandström, L. P. Rajukumar, M. Terrones, T. Wågberg and E. Gracia-Espino, Adv. Funct. Mater., 2018, 28, 1802744 CrossRef .
  88. X. Geng, W. Sun, W. Wu, B. Chen, A. Al-Hilo, M. Benamara, H. Zhu, F. Watanabe, J. Cui and T.-P. Chen, Nat. Commun., 2016, 7, 10672 CrossRef CAS PubMed .
  89. W. Chen, J. Gu, Q. Liu, R. Luo, L. Yao, B. Sun, W. Zhang, H. Su, B. Chen, P. Liu and D. Zhang, ACS Nano, 2018, 12, 308–316 CrossRef CAS PubMed .
  90. C. Tan, Z. Luo, A. Chaturvedi, Y. Cai, Y. Du, Y. Gong, Y. Huang, Z. Lai, X. Zhang, L. Zheng, X. Qi, M. H. Goh, J. Wang, S. Han, X.-J. Wu, L. Gu, C. Kloc and H. Zhang, Adv. Mater., 2018, 30, 1705509 CrossRef PubMed .
  91. J. Zhu, Z.-C. Wang, H. Dai, Q. Wang, R. Yang, H. Yu, M. Liao, J. Zhang, W. Chen, Z. Wei, N. Li, L. Du, D. Shi, W. Wang, L. Zhang, Y. Jiang and G. Zhang, Nat. Commun., 2019, 10, 1348 CrossRef PubMed .
  92. Z. Liu, K. Nie, X. Qu, X. Li, B. Li, Y. Yuan, S. Chong, P. Liu, Y. Li, Z. Yin and W. Huang, J. Am. Chem. Soc., 2022, 144, 4863–4873 CrossRef CAS PubMed .
  93. L. Ji, P. Yan, C. Zhu, C. Ma, W. Wu, C. Wei, Y. Shen, S. Chu, J. Wang, Y. Du, J. Chen, X. Yang and Q. Xu, Appl. Catal., B, 2019, 251, 87–93 CrossRef CAS .
  94. L. Liu, J. Wu, L. Wu, M. Ye, X. Liu, Q. Wang, S. Hou, P. Lu, L. Sun, J. Zheng, L. Xing, L. Gu, X. Jiang, L. Xie and L. Jiao, Nat. Mater., 2018, 17, 1108–1114 CrossRef CAS PubMed .
  95. X. Qian, J. Liu, L. Fu and J. Li, Science, 2014, 346, 1344–1347 CrossRef CAS PubMed .
  96. G. Gao, Y. Jiao, F. Ma, Y. Jiao, E. Waclawik and A. Du, J. Phys. Chem. C, 2015, 119, 13124–13128 CrossRef CAS .
  97. Y. Yu, G.-H. Nam, Q. He, X.-J. Wu, K. Zhang, Z. Yang, J. Chen, Q. Ma, M. Zhao, Z. Liu, F.-R. Ran, X. Wang, H. Li, X. Huang, B. Li, Q. Xiong, Q. Zhang, Z. Liu, L. Gu, Y. Du, W. Huang and H. Zhang, Nat. Chem., 2018, 10, 638–643 CrossRef CAS PubMed .
  98. G. H. Nam, Q. He, X. Wang, Y. Yu, J. Chen, K. Zhang, Z. Yang, D. Hu, Z. Lai, B. Li, Q. Xiong, Q. Zhang, L. Gu and H. Zhang, Adv. Mater., 2019, 31, 1807764 CrossRef PubMed .
  99. N. H. Attanayake, A. C. Thenuwara, A. Patra, Y. V. Aulin, T. M. Tran, H. Chakraborty, E. Borguet, M. L. Klein, J. P. Perdew and D. R. Strongin, ACS Energy Lett., 2017, 3, 7–13 CrossRef .
  100. T. H. Lau, S. Wu, R. Kato, T.-S. Wu, J. Kulhavy, J. Mo, J. Zheng, J. S. Foord, Y.-L. Soo, K. Suenaga, M. T. Darby and S. C. E. Tsang, ACS Catal., 2019, 9, 7527–7534 CrossRef CAS .
  101. Q. Tang and D.-e Jiang, ACS Catal., 2016, 6, 4953–4961 CrossRef CAS .
  102. Q. Jin, N. Liu, C. Dai, R. Xu, B. Wu, G. Yu, B. Chen and Y. Du, Adv. Energy Mater., 2020, 10, 2000291 CrossRef CAS .
  103. K. Qi, X. Cui, L. Gu, S. Yu, X. Fan, M. Luo, S. Xu, N. Li, L. Zheng, Q. Zhang, J. Ma, Y. Gong, F. Lv, K. Wang, H. Huang, W. Zhang, S. Guo, W. Zheng and P. Liu, Nat. Commun., 2019, 10, 5231 CrossRef PubMed .
  104. B. Pattengale, Y. Huang, X. Yan, S. Yang, S. Younan, W. Hu, Z. Li, S. Lee, X. Pan, J. Gu and J. Huang, Nat. Commun., 2020, 11, 4114 CrossRef CAS PubMed .
  105. J. Shi, P. Yu, F. Liu, P. He, R. Wang, L. Qin, J. Zhou, X. Li, J. Zhou, X. Sui, S. Zhang, Y. Zhang, Q. Zhang, T. C. Sum, X. Qiu, Z. Liu and X. Liu, Adv. Mater., 2017, 29, 1701486 CrossRef PubMed .
  106. R. Suzuki, M. Sakano, Y. Zhang, R. Akashi, D. Morikawa, A. Harasawa, K. Yaji, K. Kuroda, K. Miyamoto, T. Okuda, K. Ishizaka, R. Arita and Y. Iwasa, Nat. Nanotechnol., 2014, 9, 611–617 CrossRef CAS PubMed .
  107. R. J. Toh, Z. Sofer, J. Luxa, D. Sedmidubský and M. Pumera, Chem. Commun., 2017, 53, 3054–3057 RSC .
  108. S. Zhou, S. Wang, H. Li, W. Xu, C. Gong, J. C. Grossman and J. H. Warner, ACS Omega, 2017, 2, 3315–3324 CrossRef CAS PubMed .
  109. J. Hu, C. Zhang, L. Jiang, H. Lin, Y. An, D. Zhou, M. K. Leung and S. Yang, Joule, 2017, 1, 383–393 CrossRef CAS .
  110. Y. Ma, D. Leng, X. Zhang, J. Fu, C. Pi, Y. Zheng, B. Gao, X. Li, N. Li, P. K. Chu, Y. Luo and K. Huo, Small, 2022, 18, 2203173 CrossRef CAS PubMed .
  111. M. Liu, J.-A. Wang, W. Klysubun, G.-G. Wang, S. Sattayaporn, F. Li, Y.-W. Cai, F. Zhang, J. Yu and Y. Yang, Nat. Commun., 2021, 12, 5260 CrossRef CAS PubMed .
  112. G. Wang, G. Zhang, X. Ke, X. Chen, X. Chen, Y. Wang, G. Huang, J. Dong, S. Chu and M. Sui, Small, 2022, 18, 2107238 CrossRef CAS PubMed .
  113. J. Liu, J. Wang, B. Zhang, Y. Ruan, H. Wan, X. Ji, K. Xu, D. Zha, L. Miao and J. Jiang, J. Mater. Chem. A, 2018, 6, 2067–2072 RSC .
  114. T. Wu, E. Song, S. Zhang, M. Luo, C. Zhao, W. Zhao, J. Liu and F. Huang, Adv. Mater., 2022, 34, 2108505 CrossRef CAS PubMed .
  115. C. Gao, Z. Li, H. Wang, Y. Yang, B. Li, Z. Peng, J. Li and Z. Liu, ChemElectroChem, 2020, 7, 355–361 CrossRef CAS .
  116. S. Huang, Y. Cao, F. Yao, D. Zhang, J. Yang, S. Ye, D. Yao, Y. Liu, J. Li, D. Lei, X. Wang, H. Huang and M. Wu, Small, 2023, 2207919 CrossRef CAS PubMed .
  117. C. Lang, W. Jiang, C. J. Yang, H. Zhong, P. Chen, Q. Wu, X. Yan, C. L. Dong, Y. Lin, L. Ouyang, Y. Jia and X. Yao, Small, 2023, 2300807 CrossRef PubMed .
  118. F. Xi, P. Bogdanoff, K. Harbauer, P. Plate, C. Höhn, J. R. Rappich, B. Wang, X. Han, R. Van De Krol and S. Fiechter, ACS Catal., 2019, 9, 2368–2380 CrossRef CAS .
  119. L. Wu, A. Longo, N. Y. Dzade, A. Sharma, M. M. Hendrix, A. A. Bol, N. H. De Leeuw, E. J. Hensen and J. P. Hofmann, ChemSusChem, 2019, 12, 4383 CrossRef CAS PubMed .
  120. S. C. Lee, J. D. Benck, C. Tsai, J. Park, A. L. Koh, F. Abild-Pedersen, T. F. Jaramillo and R. Sinclair, ACS Nano, 2016, 10, 624–632 CrossRef CAS PubMed .
  121. B. Lassalle-Kaiser, D. Merki, H. Vrubel, S. Gul, V. K. Yachandra, X. Hu and J. Yano, J. Am. Chem. Soc., 2015, 137, 314–321 CrossRef CAS PubMed .
  122. P. D. Tran, T. V. Tran, M. Orio, S. Torelli, Q. D. Truong, K. Nayuki, Y. Sasaki, S. Y. Chiam, R. Yi, I. Honma, J. Barber and V. Artero, Nat. Mater., 2016, 15, 640–646 CrossRef CAS PubMed .
  123. Y. Deng, L. R. L. Ting, P. H. L. Neo, Y.-J. Zhang, A. A. Peterson and B. S. Yeo, ACS Catal., 2016, 6, 7790–7798 CrossRef CAS .
  124. L. R. L. Ting, Y. Deng, L. Ma, Y.-J. Zhang, A. A. Peterson and B. S. Yeo, ACS Catal., 2016, 6, 861–867 CrossRef CAS .
  125. B. Li, L. Jiang, X. Li, Z. Cheng, P. Ran, P. Zuo, L. Qu, J. Zhang and Y. Lu, Adv. Funct. Mater., 2019, 29, 1806229 CrossRef .
  126. Y. Li, Y. Yu, Y. Huang, R. A. Nielsen, W. A. Goddard III, Y. Li and L. Cao, ACS Catal., 2015, 5, 448–455 CrossRef CAS .
  127. R. Ding, M. Wang, X. Wang, H. Wang, L. Wang, Y. Mu and B. Lv, Nanoscale, 2019, 11, 11217–11226 RSC .
  128. D. Zhang, F. Wang, W. Zhao, M. Cui, X. Fan, R. Liang, Q. Ou and S. Zhang, Adv. Sci., 2022, 9, 2202445 CrossRef CAS PubMed .
  129. H. Vrubel and X. Hu, ACS Catal., 2013, 3, 2002–2011 CrossRef CAS .
  130. T. R. Hellstern, J. Kibsgaard, C. Tsai, D. W. Palm, L. A. King, F. Abild-Pedersen and T. F. Jaramillo, ACS Catal., 2017, 7, 7126–7130 CrossRef CAS .
  131. P. K. Holzapfel, M. Bühler, D. Escalera-López, M. Bierling, F. D. Speck, K. J. Mayrhofer, S. Cherevko, C. V. Pham and S. Thiele, Small, 2020, 16, 2003161 CrossRef CAS PubMed .
  132. B. Seo, G. Y. Jung, S. J. Lee, D. S. Baek, Y. J. Sa, H. W. Ban, J. S. Son, K. Park, S. K. Kwak and S. H. Joo, ACS Catal., 2019, 10, 652–662 CrossRef .
  133. T. F. Jaramillo, J. Bonde, J. Zhang, B.-L. Ooi, K. Andersson, J. Ulstrup and I. Chorkendorff, J. Phys. Chem. C, 2008, 112, 17492–17498 CrossRef CAS .
  134. Z. Huang, W. Luo, L. Ma, M. Yu, X. Ren, M. He, S. Polen, K. Click, B. Garrett, J. Lu, K. Amine, C. Hadad, W. Chen, A. Asthagiri and Y. Wu, Angew. Chem., Int. Ed., 2015, 54, 15181–15185 CrossRef CAS PubMed .
  135. C.-H. Lee, S. Lee, Y.-K. Lee, Y. C. Jung, Y.-I. Ko, D. C. Lee and H.-I. Joh, ACS Catal., 2018, 8, 5221–5227 CrossRef CAS .
  136. J. McAllister, N. A. Bandeira, J. C. McGlynn, A. Y. Ganin, Y.-F. Song, C. Bo and H. N. Miras, Nat. Commun., 2019, 10, 370 CrossRef CAS PubMed .
  137. Z. Ji, C. Trickett, X. Pei and O. M. Yaghi, J. Am. Chem. Soc., 2018, 140, 13618–13622 CrossRef CAS PubMed .
  138. B. R. Garrett, S. M. Polen, M. Pimplikar, C. M. Hadad and Y. Wu, J. Am. Chem. Soc., 2017, 139, 4342–4345 CrossRef CAS PubMed .
  139. B. R. Garrett, K. A. Click, C. B. Durr, C. M. Hadad and Y. Wu, J. Am. Chem. Soc., 2016, 138, 13726–13731 CrossRef CAS PubMed .
  140. Y. He, Q. He, L. Wang, C. Zhu, P. Golani, A. D. Handoko, X. Yu, C. Gao, M. Ding, X. Wang, F. Liu, Q. Zeng, P. Yu, S. Guo, B. I. Yakobson, L. Wang, Z. W. Seh, Z. Zhang, M. Wu, Q. J. Wang, H. Zhang and Z. Liu, Nat. Mater., 2019, 18, 1098–1104 CrossRef CAS PubMed .
  141. J. Zhang, X. Tian, M. Liu, H. Guo, J. Zhou, Q. Fang, Z. Liu, Q. Wu and J. Lou, J. Am. Chem. Soc., 2019, 141, 19269–19275 CrossRef CAS PubMed .
  142. X. Li, T. Li, Y. Ma, Q. Wei, W. Qiu, H. Guo, X. Shi, P. Zhang, A. M. Asiri, L. Chen, B. Tang and X. Sun, Adv. Energy Mater., 2018, 8, 1801357 CrossRef .
  143. J. Chen, C. Zhang, M. Huang, J. Zhang, J. Zhang, H. Liu, G. Wang and R. Wang, Appl. Catal., B, 2021, 285, 119810 CrossRef CAS .
  144. S. B. Patil, H.-L. Chou, Y.-M. Chen, S.-H. Hsieh, C.-H. Chen, C.-C. Chang, S.-R. Li, Y.-C. Lee, Y.-S. Lin, H. Li, Y. J. Chang, Y.-H. Lai and D.-Y. Wang, J. Mater. Chem. A, 2021, 9, 1230–1239 RSC .
  145. W. Wu, C. Niu, P. Yan, F. Shi, C. Ma, X. Yang, Y. Jia, J. Chen, M. I. Ahmed, C. Zhao and Q. Xu, Appl. Catal., B, 2021, 298, 120615 CrossRef CAS .
  146. R. Liu, T. Guo, H. Fei, Z. Wu, D. Wang and F. Liu, Adv. Sci., 2022, 9, 2103583 CrossRef CAS PubMed .
  147. G. Lin, Q. Ju, X. Guo, W. Zhao, S. Adimi, J. Ye, Q. Bi, J. Wang, M. Yang and F. Huang, Adv. Mater., 2021, 33, 2007509 CrossRef CAS PubMed .
  148. J. Liang, S. Ma, J. Li, Y. Wang, J. Wu, Q. Zhang, Z. Liu, Z. Yang, K. Qu and W. Cai, J. Mater. Chem. A, 2020, 8, 10426–10432 RSC .
  149. J. Zhao, J. Zhao and Q. Cai, Phys. Chem. Chem. Phys., 2018, 20, 9248–9255 RSC .
  150. T. Yang, T. T. Song, J. Zhou, S. Wang, D. Chi, L. Shen, M. Yang and Y. P. Feng, Nano Energy, 2020, 68, 104304 CrossRef CAS .
  151. Q. Dang, S. Tang, T. Liu, X. Li, X. Wang, W. Zhong, Y. Luo and J. Jiang, J. Phys. Chem. Lett., 2021, 12, 8355–8362 CrossRef CAS PubMed .
  152. L. Chen, Y. Zhou, Y. Rao, M. Qin, B. Gong, W. Zhang and Z. Li, J. Phys. Chem. C, 2022, 126, 9352–9360 CrossRef CAS .
  153. K. Xie, F. Wang, F. Wei, J. Zhao and S. Lin, J. Phys. Chem. C, 2022, 126, 5180–5188 CrossRef CAS .
  154. J. Li, S. Chen, F. Quan, G. Zhan, F. Jia, Z. Ai and L. Zhang, Chem, 2020, 6, 885–901 CAS .
  155. H. Su, L. Chen, Y. Chen, R. Si, Y. Wu, X. Wu, Z. Geng, W. Zhang and J. Zeng, Angew. Chem., Int. Ed., 2020, 59, 20411–20416 CrossRef CAS PubMed .
  156. Z. He, Y. Jiang, X. Cui, Z. Liu, X. Meng, J. Wan and F. Ma, ACS Appl. Nano Mater., 2022, 5, 5470–5478 CrossRef CAS .
  157. H. Fei, T. Guo, Y. Xin, L. Wang, R. Liu, D. Wang, F. Liu and Z. Wu, Appl. Catal., B, 2022, 300, 120733 CrossRef CAS .
  158. X. Chen, C. Ma, Z. Tan, X. Wang, X. Qian, X. Zhang, J. Tian, S. Yan and M. Shao, Chem. Eng. J., 2022, 433, 134504 CrossRef CAS .
  159. X. Chen, S. Zhang, X. Qian, Z. Liang, Y. Xue, X. Zhang, J. Tian, Y. Han and M. Shao, Appl. Catal., B, 2022, 310, 121277 CrossRef CAS .
  160. X. Xu, X. Tian, B. Sun, Z. Liang, H. Cui, J. Tian and M. Shao, Appl. Catal., B, 2020, 272, 118984 CrossRef CAS .
  161. X. Zi, J. Wan, X. Yang, W. Tian, H. Zhang and Y. Wang, Appl. Catal., B, 2021, 286, 119870 CrossRef CAS .
  162. B. H. Suryanto, D. Wang, L. M. Azofra, M. Harb, L. Cavallo, R. Jalili, D. R. Mitchell, M. Chatti and D. R. MacFarlane, ACS Energy Lett., 2018, 4, 430–435 CrossRef .
  163. W. Zhang, Y. Hu, L. Ma, G. Zhu, Y. Wang, X. Xue, R. Chen, S. Yang and Z. Jin, Adv. Sci., 2018, 5, 1700275 CrossRef PubMed .
  164. M. Asadi, K. Kim, C. Liu, A. V. Addepalli, P. Abbasi, P. Yasaei, P. Phillips, A. Behranginia, J. M. Cerrato, R. Haasch, P. Zapol, B. Kumar, R. F. Klie, J. Abiade, L. A. Curtiss and A. Salehi-Khojin, Science, 2016, 353, 467–470 CrossRef CAS PubMed .
  165. P. Abbasi, M. Asadi, C. Liu, S. Sharifi-Asl, B. Sayahpour, A. Behranginia, P. Zapol, R. Shahbazian-Yassar, L. A. Curtiss and A. Salehi-Khojin, ACS Nano, 2017, 11, 453–460 CrossRef CAS PubMed .
  166. K. Lv, W. Suo, M. Shao, Y. Zhu, X. Wang, J. Feng, M. Fang and Y. Zhu, Nano Energy, 2019, 63, 103834 CrossRef CAS .
  167. S. A. Francis, J. M. Velazquez, I. M. Ferrer, D. A. Torelli, D. Guevarra, M. T. McDowell, K. Sun, X. Zhou, F. H. Saadi, J. John, M. H. Richter, F. P. Hyler, K. M. Papadantonakis, B. S. Brunschwig and N. S. Lewis, Chem. Mater., 2018, 30, 4902–4908 CrossRef CAS .
  168. X. Mao, L. Wang, Y. Xu and Y. Li, J. Phys. Chem. C, 2020, 124, 10523–10529 CrossRef CAS .
  169. W. Huang, D. Zhou, H. Yang, X. Liu and J. Luo, ACS Appl. Energy Mater., 2021, 4, 7492–7496 CrossRef CAS .
  170. J. Xu, X. Li, W. Liu, Y. Sun, Z. Ju, T. Yao, C. Wang, H. Ju, J. Zhu, S. Wei and Y. Xie, Angew. Chem., Int. Ed., 2017, 56, 9121–9125 CrossRef CAS PubMed .
  171. H. Li, X. Liu, S. Chen, D. Yang, Q. Zhang, L. Song, H. Xiao, Q. Zhang, L. Gu and X. Wang, Adv. Energy Mater., 2019, 9, 1900072 CrossRef .
  172. K. Lv, C. Teng, M. Shi, Y. Yuan, Y. Zhu, J. Wang, Z. Kong, X. Lu and Y. Zhu, Adv. Funct. Mater., 2018, 28, 1802339 CrossRef .
  173. F. Li, S.-F. Zhao, L. Chen, A. Khan, D. R. MacFarlane and J. Zhang, Energy Environ. Sci., 2016, 9, 216–223 RSC .
  174. Y. Linghu, T. Tong, C. Li and C. Wu, Appl. Surf. Sci., 2022, 590, 153001 CrossRef CAS .
  175. Y. Ren, X. Sun, K. Qi and Z. Zhao, Appl. Surf. Sci., 2022, 154211 CrossRef CAS .
  176. Z.-Z. Lin, X.-M. Li, X.-W. Chen and X. Chen, Phys. Chem. Chem. Phys., 2022, 24, 3733–3740 RSC .
  177. Z. W. Chen, W. Gao, W. T. Zheng and Q. Jiang, ChemSusChem, 2018, 11, 1455–1459 CrossRef CAS PubMed .
  178. S. Kang, S. Han and Y. Kang, ChemSusChem, 2019, 12, 2671–2678 CrossRef CAS PubMed .
  179. X. Hong, K. Chan, C. Tsai and J. K. Nørskov, ACS Catal., 2016, 6, 4428–4437 CrossRef CAS .
  180. X. Yao, Z. W. Chen, G. J. Liu, X. Y. Lang, Y. F. Zhu, W. Gao and Q. Jiang, ChemSusChem, 2021, 14, 2255–2261 CrossRef CAS PubMed .
  181. P. Li, H. Hu, J. Xu, H. Jing, H. Peng, J. Lu, C. Wu and S. Ai, Appl. Catal., B, 2014, 147, 912–919 CrossRef CAS .
  182. H. Peng, J. Lu, C. Wu, Z. Yang, H. Chen, W. Song, P. Li and H. Yin, Appl. Surf. Sci., 2015, 353, 1003–1012 CrossRef CAS .
  183. Y. Wang, Z. Zhang, L. Zhang, Z. Luo, J. Shen, H. Lin, J. Long, J. C. Wu, X. Fu, X. Wang and C. Li, J. Am. Chem. Soc., 2018, 140, 14595–14598 CrossRef CAS PubMed .
  184. F. Yu, X. Jing, Y. Wang, M. Sun and C. Duan, Angew. Chem., Int. Ed., 2021, 60, 24849–24853 CrossRef CAS PubMed .
  185. S. F. Ng, M. Y. L. Lau and W. J. Ong, Adv. Mater., 2021, 33, 2008654 CrossRef CAS PubMed .
  186. M. Wang, H. Yang, K. Shen, H. Xu, W. Wang, Z. Yang, L. Zhang, J. Chen, Y. Huang, M. Chen, D. Mitlin and X. Li, Small Methods, 2020, 4, 2000353 CrossRef CAS .
  187. H. Lin, S. Zhang, T. Zhang, H. Ye, Q. Yao, G. W. Zheng and J. Y. Lee, Adv. Energy Mater., 2019, 9, 1902096 CrossRef CAS .
  188. C. Li, Z. Xi, D. Guo, X. Chen and L. Yin, Small, 2018, 14, 1701986 CrossRef PubMed .
  189. S. Rehman, K. Khan, Y. Zhao and Y. Hou, J. Mater. Chem. A, 2017, 5, 3014–3038 RSC .
  190. J. Yan, X. Liu and B. Li, Adv. Sci., 2016, 3, 1600101 CrossRef PubMed .
  191. B. Li, Q. Su, L. Yu, J. Zhang, G. Du, D. Wang, D. Han, M. Zhang, S. Ding and B. Xu, ACS Nano, 2020, 14, 17285–17294 CrossRef CAS PubMed .
  192. D. Liu, C. Zhang, G. Zhou, W. Lv, G. Ling, L. Zhi and Q. H. Yang, Adv. Sci., 2018, 5, 1700270 CrossRef PubMed .
  193. S. Luo, J. Ruan, Y. Wang, J. Hu, Y. Song, M. Chen and L. Wu, Small, 2021, 17, 2101879 CrossRef CAS PubMed .
  194. Z. Ma, Y. Liu, J. Gautam, W. Liu, A. N. Chishti, J. Gu, G. Yang, Z. Wu, J. Xie, M. Chen, L. Ni and G. Diao, Small, 2021, 17, 2102710 CrossRef CAS PubMed .
  195. T. Yang, B. Guo, W. Du, M. K. Aslam, M. Tao, W. Zhong, Y. Chen, S. J. Bao, X. Zhang and M. Xu, Adv. Sci., 2019, 6, 1901557 CrossRef CAS PubMed .
  196. H. Wang, Q. Zhang, H. Yao, Z. Liang, H.-W. Lee, P.-C. Hsu, G. Zheng and Y. Cui, Nano Lett., 2014, 14, 7138–7144 CrossRef CAS PubMed .
  197. G. Babu, N. Masurkar, H. Al Salem and L. M. R. Arava, J. Am. Chem. Soc., 2017, 139, 171–178 CrossRef CAS PubMed .
  198. H. Lin, L. Yang, X. Jiang, G. Li, T. Zhang, Q. Yao, G. W. Zheng and J. Y. Lee, Energy Environ. Sci., 2017, 10, 1476–1486 RSC .
  199. Y. Pan, L. Gong, X. Cheng, Y. Zhou, Y. Fu, J. Feng, H. Ahmed and H. Zhang, ACS Nano, 2020, 14, 5917–5925 CrossRef CAS PubMed .
  200. Z. Du, Y. Guo, H. Wang, J. Gu, Y. Zhang, Z. Cheng, B. Li, S. Li and S. Yang, ACS Nano, 2021, 15, 19275–19283 CrossRef CAS PubMed .
  201. Y. Zhang, Z. Mu, C. Yang, Z. Xu, S. Zhang, X. Zhang, Y. Li, J. Lai, Z. Sun, Y. Yang, Y. Chao, C. Li, X. Ge, W. Yang and S. Guo, Adv. Funct. Mater., 2018, 28, 1707578 CrossRef .
  202. Z. Cheng, Y. Chen, Y. Yang, L. Zhang, H. Pan, X. Fan, S. Xiang and Z. Zhang, Adv. Energy Mater., 2021, 11, 2003718 CrossRef CAS .
  203. Z.-L. Xu, N. Onofrio and J. Wang, J. Mater. Chem. A, 2020, 8, 17646–17656 RSC .
  204. C. Tian, B. Li, X. Hu, J. Wu, P. Li, X. Xiang, X. Zu and S. Li, ACS Appl. Mater. Interfaces, 2021, 13, 6229–6240 CrossRef CAS PubMed .
  205. W. Liu, C. Luo, S. Zhang, B. Zhang, J. Ma, X. Wang, W. Liu, Z. Li, Q.-H. Yang and W. Lv, ACS Nano, 2021, 15, 7491–7499 CrossRef CAS PubMed .
  206. Y. Li, M. Chen, H. Liu, P. Zeng, D. Zhang, H. Yu, X. Zhou, C. Guo, Z. Luo, Y. Wang, B. Chang and X. Wang, Chem. Eng. J., 2021, 425, 131640 CrossRef CAS .
  207. S. Bhoyate, J. Kim, E. Lee, B. Park, E. Lee, J. Park, S. H. Oh, J. Kim and W. Choi, J. Mater. Chem. A, 2020, 8, 12436–12445 RSC .
  208. S. Bhoyate, B. Park, S. H. Oh and W. Choi, J. Power Sources, 2021, 511, 230426 CrossRef CAS .
  209. K. Liu, X. Zhang, F. Miao, Z. Wang, S. Zhang, Y. Zhang, P. Zhang and G. Shao, Small, 2021, 17, 2100065 CrossRef CAS PubMed .
  210. H. Song, T. Li, T. He, Z. Wang, D. Fang, Y. Wang, X. L. Li, D. Zhang, J. Hu and S. Huang, Chem. Eng. J., 2022, 450, 138115 CrossRef CAS .
  211. Y. Li, X. Yan, Z. Zhou, J. Liu, Z. Zhang, X. Guo, H. Peng and G. Li, Appl. Surf. Sci., 2022, 574, 151586 CrossRef CAS .
  212. S. Wang, S. Feng, J. Liang, Q. Su, F. Zhao, H. Song, M. Zheng, Q. Sun, Z. Song, X. Jia, J. Yang, Y. Li, J. Liao, R. Li and X. Sun, Adv. Energy Mater., 2021, 11, 2003314 CrossRef CAS .
  213. J. He, G. Hartmann, M. Lee, G. S. Hwang, Y. Chen and A. Manthiram, Energy Environ. Sci., 2019, 12, 344–350 RSC .
  214. Q. Wu, Z. Yao, X. Zhou, J. Xu, F. Cao and C. Li, ACS Nano, 2020, 14, 3365–3377 CrossRef CAS PubMed .
  215. M. Wang, L. Fan, D. Tian, X. Wu, Y. Qiu, C. Zhao, B. Guan, Y. Wang, N. Zhang and K. Sun, ACS Energy Lett., 2018, 3, 1627–1633 CrossRef CAS .
  216. M. Song, H. Tan, X. Li, A. I. Y. Tok, P. Liang, D. Chao and H. J. Fan, Small Methods, 2020, 4, 1900274 CrossRef CAS .
  217. W. Wang and Y.-C. Lu, Acc. Mater. Res., 2021, 2, 515–525 CrossRef CAS .
  218. Y. Yan, S. Liang, X. Wang, M. Zhang, S.-M. Hao, X. Cui, Z. Li and Z. Lin, Proc. Natl. Acad. Sci. U. S. A., 2021, 118, e2110036118 CrossRef CAS PubMed .
  219. Z. Sadighi, J. Liu, L. Zhao, F. Ciucci and J.-K. Kim, Nanoscale, 2018, 10, 22549–22559 RSC .
  220. Y. Liu, Y. Zang, X. Liu, J. Cai, Z. Lu, S. Niu, Z. Pei, T. Zhai and G. Wang, Front. Energy Res., 2020, 8, 109 CrossRef .
  221. X. Fan, Y. Huang, H. Wang, F. Zheng, C. Tan, Y. Li, X. Lu, Z. Ma and Q. Li, Electrochim. Acta, 2021, 369, 137653 CrossRef CAS .
  222. S. Zhang, Z. Huang, Z. Wen, L. Zhang, J. Jin, R. Shahbazian-Yassar and J. Yang, Nano Lett., 2017, 17, 3518–3526 CrossRef CAS PubMed .
  223. L. Li, C. Chen, J. Su, P. Kuang, C. Zhang, Y. Yao, T. Huang and A. Yu, J. Mater. Chem. A, 2016, 4, 10986–10991 RSC .
  224. S. Wu, W. Di, D. Zhang, W. Liu, H. Luo, J. He, Q. Yang, Z. Li and R. Liu, Energy Environ. Mater., 2022, 5, 918–927 CrossRef CAS .
  225. A. Jaradat, C. Zhang, S. K. Singh, J. Ahmed, A. Ahmadiparidari, L. Majidi, S. Rastegar, Z. Hemmat, S. Wang, A. T. Ngo, L. A. Curtiss, M. Daly, A. Subramanian and A. Salehi-khojin, Small, 2021, 17, 2102072 CrossRef CAS PubMed .
  226. M. Asadi, B. Sayahpour, P. Abbasi, A. T. Ngo, K. Karis, J. R. Jokisaari, C. Liu, B. Narayanan, M. Gerard, P. Yasaei, X. Hu, A. Mukherjee, K. C. Lau, R. S. Assary, F. Khalili-Araghi, R. F. Klie, L. A. Curtiss and A. Salehi-Khojin, Nature, 2018, 555, 502–506 CrossRef CAS PubMed .
  227. A. Hu, J. Long, C. Shu, R. Liang and J. Li, ACS Appl. Mater. Interfaces, 2018, 10, 34077–34086 CrossRef CAS PubMed .
  228. A. Hu, C. Shu, X. Qiu, M. Li, R. Zheng and J. Long, ACS Sustainable Chem. Eng., 2019, 7, 6929–6938 CrossRef CAS .
  229. D. Li, L. Zhao, Q. Xia, J. Wang, X. Liu, H. Xu and S. Chou, Adv. Funct. Mater., 2022, 32, 2108153 CrossRef CAS .
  230. D. Li, L. Zhao, Q. Xia, L. Liu, W. Fang, Y. Liu, Z. Zhou, Y. Long, X. Han, Y. Zhang, J. Wang, Y. Wu and H. Liu, Small, 2022, 18, 2105752 CrossRef CAS PubMed .
  231. X. Wen, D. Du, L. Ren, H. Xu, R. Li, C. Zhao and C. Shu, Chem. Eng. J., 2022, 442, 136311 CrossRef CAS .
  232. Z. Sun, J. He, M. Yuan, L. Lin, Z. Zhang, Z. Kang, Q. Liao, H. Li, G. Sun, X. Yang, R. Long and Y. Zhang, Nano Energy, 2019, 65, 103996 CrossRef CAS .
  233. G. Solomon, M. G. Kohan, M. Vagin, F. Rigoni, R. Mazzaro, M. M. Natile, S. You, V. Morandi, I. Concina and A. Vomiero, Nano Energy, 2021, 81, 105664 CrossRef CAS .
  234. P. Zhang, X. Lu, Y. Huang, J. Deng, L. Zhang, F. Ding, Z. Su, G. Wei and O. G. Schmidt, J. Mater. Chem. A, 2015, 3, 14562–14566 RSC .
  235. X. Wang, Y. Li, X. Bi, L. Ma, T. Wu, M. Sina, S. Wang, M. Zhang, J. Alvarado, B. Lu, A. Banerjee, K. Amine, J. Lu and Y. S. Meng, Joule, 2018, 2, 2381–2392 CrossRef CAS .
  236. I. S. Amiinu, Z. Pu, X. Liu, K. A. Owusu, H. G. R. Monestel, F. O. Boakye, H. Zhang and S. Mu, Adv. Funct. Mater., 2017, 27, 1702300 CrossRef .
  237. M. K. Puglia, M. Malhotra, A. Chivukula and C. V. Kumar, ACS Appl. Nano Mater., 2021, 4, 10389–10398 CrossRef CAS .
  238. C. He, L. Cui, X. Wu, Y. Cai, Z. Ma, X. Zhong, H. Wang, Q. Li, Y. Huang and Q. Pan, ACS Appl. Energy Mater., 2021, 4, 9498–9506 CrossRef CAS .
  239. P. Basumatary, D. Konwar and Y. S. Yoon, Appl. Catal., B, 2020, 267, 118724 CrossRef CAS .
  240. A. Ahmadiparidari, R. E. Warburton, L. Majidi, M. Asadi, A. Chamaani, J. R. Jokisaari, S. Rastegar, Z. Hemmat, B. Sayahpour, R. S. Assary, B. Narayanan, P. Abbasi, P. C. Redfern, A. Ngo, M. Vörös, J. Greeley, R. Klie, L. A. Curtiss and A. Salehi-Khojin, Adv. Mater., 2019, 31, 1902518 CrossRef CAS PubMed .
  241. C.-J. Chen, C.-S. Huang, Y.-C. Huang, F.-M. Wang, X.-C. Wang, C.-C. Wu, W.-S. Chang, C.-L. Dong, L.-C. Yin and R.-S. Liu, ACS Appl. Mater. Interfaces, 2021, 13, 6156–6167 CrossRef CAS PubMed .
  242. N. Salazar, S. Rangarajan, J. Rodríguez-Fernández, M. Mavrikakis and J. V. Lauritsen, Nat. Commun., 2020, 11, 4369 CrossRef PubMed .
  243. P. Zheng, A. Duan, K. Chi, L. Zhao, C. Zhang, C. Xu, Z. Zhao, W. Song, X. Wang and J. Fan, Chem. Eng. Sci., 2017, 164, 292–306 CrossRef CAS .
  244. P. Nikulshin, V. Salnikov, A. Mozhaev, P. Minaev, V. Kogan and A. Pimerzin, J. Catal., 2014, 309, 386–396 CrossRef CAS .
  245. S. P. Lonkar, V. V. Pillai and S. M. Alhassan, ACS Appl. Nano Mater., 2018, 1, 3114–3123 CrossRef CAS .
  246. X.-z Wang, N. Liu, H. Hu, X.-p Wang and J.-s Qiu, Carbon, 2014, 76, 471 CrossRef .
  247. C. F. Weise, H. Falsig, P. G. Moses, S. Helveg, M. Brorson and L. P. Hansen, J. Catal., 2021, 403, 74–86 CrossRef CAS .
  248. S. S. Grønborg, N. Salazar, A. Bruix, J. Rodríguez-Fernández, S. D. Thomsen, B. Hammer and J. V. Lauritsen, Nat. Commun., 2018, 9, 2211 CrossRef PubMed .
  249. H. Shang, T. Wang and W. Zhang, Chem. Eng. Sci., 2019, 195, 208–217 CrossRef CAS .
  250. R. V. Mom, J. N. Louwen, J. W. Frenken and I. M. Groot, Nat. Commun., 2019, 10, 2546 CrossRef PubMed .
  251. A. Tuxen, J. Kibsgaard, H. Gøbel, E. Lægsgaard, H. Topsøe, J. V. Lauritsen and F. Besenbacher, ACS Nano, 2010, 4, 4677–4682 CrossRef CAS PubMed .
  252. J.-F. Paul and E. Payen, J. Phys. Chem. B, 2003, 107, 4057–4064 CrossRef CAS .
  253. G. Liu, A. W. Robertson, M. M.-J. Li, W. C. Kuo, M. T. Darby, M. H. Muhieddine, Y.-C. Lin, K. Suenaga, M. Stamatakis, J. H. Warner and S. C. E. Tsang, Nat. Chem., 2017, 9, 810–816 CrossRef CAS PubMed .
  254. K. Wu, W. Wang, H. Guo, Y. Yang, Y. Huang, W. Li and C. Li, ACS Energy Lett., 2020, 5, 1330–1336 CrossRef CAS .
  255. X. Bai, Q. Li, L. Shi, C. Ling and J. Wang, Catal. Today, 2020, 350, 56–63 CrossRef CAS .
  256. J. Hu, L. Yu, J. Deng, Y. Wang, K. Cheng, C. Ma, Q. Zhang, W. Wen, S. Yu, Y. Pan, J. Yang, H. Ma, F. Qi, Y. Wang, Y. Zheng, M. Chen, R. Huang, S. Zhang, Z. Zhao, J. Mao, X. Meng, Q. Ji, G. Hou, X. Han, X. Bao, Y. Wang and D. Deng, Nat. Catal., 2021, 4, 242–250 CrossRef CAS .

This journal is © The Royal Society of Chemistry 2023
Click here to see how this site uses Cookies. View our privacy policy here.