Haiyan
Xiang
a,
Qizhi
Dong
*b,
Meiqing
Yang
c and
Song
Liu
*b
aCollege of Materials and Advanced Manufacturing, Hunan University of Technology, Zhuzhou 412007, China
bState Key Laboratory of Chemo/Biosensing and Chemometrics, College of Chemistry and Chemical Engineering, Hunan University, Changsha 410082, China. E-mail: lili2sohu@hnu.edu.cn; liusong@hnu.edu.cn
cCollege of Life and Environmental Science, Hunan University of Arts and Science, Changde 415000, China
First published on 7th February 2024
Transition metal selenides (TMSes) have gained significant attention as viable non-noble metal catalysts for water splitting due to their low cost and potential electrocatalytic activity. However, the electrocatalytic activity of TMSes can be hindered by poor conductivity and limited active sites. To overcome these limitations, numerous TMSe-based electrocatalysts have been recently designed and reported, aiming to optimize their performance. This review provides a comprehensive summary of the recent progress in TMSe-based electrocatalysts for hydrogen evolution reaction (HER), oxygen evolution reaction (OER), and overall water splitting. Additionally, we discuss the various preparation methods and types of TMSe-based electrocatalysts. Furthermore, we outline future perspectives for the advancement of TMSe-based electrocatalysts. Overall, this review aims to offer new insights into the rational design and application of TMSe-based electrocatalysts for efficient water splitting.
Nanostructured transition metal dichalcogenides (TMDs) have been proven to be exceptional catalysts for both OER and HER, demonstrating superior performance and high robustness.11–13 The emerging transition metal selenides (TMSes) share a similar structure to their corresponding TMDs. Selenium (Se) atoms possess an electronic structure of 4s24p4 and can readily gain two electrons from less electronegative elements to form Se2− ions or share electrons with more electronegative elements to form TMSes. TMSes can be primarily classified into two typical types based on their morphology: graphite-like layered MSe2 structures (M = Mo, W, etc.) and non-layered MxSey structures (M = Co, Ni, Fe, etc.). Each monolayer of MSe2 consists of three atomic layers that are bonded by strong covalent bonds and linked to another layer through weak van der Waals forces. MSe2 exhibits three well-known phases: 1T with octahedral coordination, 2H with trigonal prismatic coordination, and 3R with trigonal prismatic coordination.14–16
TMSes generally exhibit higher electrical conductivity. For example, MoSe2 demonstrates higher intrinsic electrical conductivity compared to MoS2 due to the more metallic nature of selenium. Therefore, TMSes are expected to serve as efficient electrocatalysts for water splitting, involving both HER and OER.17,18 The first transition metal selenides (NiSe) were prepared in 2012,19 and were subsequently utilized as catalysts for HER in acidic media. In 2013, MoSe2 and WSe2 were synthesized and proved to be efficient catalysts for electrocatalytic HER.20 Since then, efforts have been made to enhance the performance of TMSes by improving the density of active sites and increasing conductivity. Various strategies have been explored, including defect engineering, control of structure and morphology, heteroatom doping, utilization of electro-conductive carbon materials as substrates, and fabrication of dual-native vacancies.21–27 Such synthetic study aims to further expand the activity for the TMSes catalysis.
In this review, we provide an overview of the recent progress in the development of metal selenides as electrocatalysts for water splitting. Firstly, we summarize the mechanism of water splitting and relevant synthetic methods, including liquid phase exfoliation (LPE), chemical vapor deposition (CVD), hydrothermal/solvothermal techniques, and others. Secondly, we discuss the strategies for improving the electrocatalytic performance of metal selenides. Finally, we address future challenges and provide perspectives on this research area. We hope that this review will offer valuable information and guidance for future studies in the field of electrocatalysis and its applications in other fields.
![]() | ||
Fig. 1 (a) Schematic diagram of an electrocatalytic water decomposition cell and HER and OER reactions.28 Reproduced with permission from ref. 28. Copyright 2021, American Chemical Society. (b) HER volcano plot for metals and MoS2. (c) OER volcano plot for metal oxides.33 Reproduced with permission from ref. 33. Copyright 2017, The American Association for the Advancement of Science. |
Volmer reaction: H3O+ + M + e− → M–H* + H2O | (1) |
Heyrovsky reaction: M–H* + H3O+ + e− → H2 + H2O + M | (2) |
Tafel reaction: M–H* + M–H*→ H2 + 2M | (3) |
In acidic and basic media, as shown in Fig. 2(a) and (b),31 a water molecule is absorbed and dissociated into H+ and OH− at the active sites; next, as in acidic solution, the Tafel reaction and Heyrovsky reaction of H ads produce the H2 molecule.32 The reaction can be expressed as below:
Volmer reaction: H2O + M + e− → M–H* + OH− | (4) |
Heyrovsky reaction: M–H* + H2O + e− → H2 + M + OH− | (5) |
Tafel reaction: M–H* + M–H* → H2 | (6) |
![]() | ||
Fig. 2 Mechanism of HER electrocatalysis in two different media. (a) Acid media. (b) Basic media. (c) The OER mechanism under alkaline conditions. (d) The typical HER and OER polarization curves.31 Reproduced with permission from ref. 31. Copyright 2022, Elsevier B.V. All rights reserved. |
The process of reaction kinetics of HER in an acid electrolyte is generally more likely to happen than that in basic and neutral environments, because of the weaker covalent bond of hydroxonium ions compared to the covalent bond of water.
Furthermore, the value of the Tafel slope can reflect the speed control step of HER. When the Tafel slope value is high, hydrogen adsorption is the speed control step. If the dynamic process of hydrogen adsorption is fast, the Tafel slope value is small, the chemical or electrochemical desorption step is the speed control step. According to the Sabatier principle, the binding energy of the catalyst and reaction intermediate should be moderate. If the hydrogen bond to the surface is too weak, adsorption (Volmer step) will limit the total reaction rate, while if the binding is too strong, the desorption (Heyrovsky/Tafel) step will limit the reaction rate. Therefore, a necessary and insufficient condition for an active catalyst is the value of ΔGH* close to zero. How to control the adsorption energy of reaction intermediates on the surface is the key to designing materials with better performance. Until now, the HER performance has been evaluated by some technical terms, e.g., overpotential, exchange current density, Tafel slope, turnover frequency, Faraday efficiency and stability. In addition, charge transfer resistance (Rct), specific surface areas and free energy differential of hydrogen adsorption (ΔGH*) are frequently used to determine the HER performance (Fig. 1(b)).33
In acidic solution
H2O + M → M–OH* + H+ + e− | (7) |
M–OH* → M–O* + H+ + e− | (8) |
2M–O* → 2M + O2 | (9) |
M–O* + H2O → M–OOH* + H+ + e− | (10) |
M–OOH* → M + O2 + H+ + e− | (11) |
In basic solution
OH− + M → M–OH* + e− | (12) |
M–OH* + OH− → M–O* + H2O + e− | (13) |
2M–O* → 2M + O2 | (14) |
M–O* + OH− → M–OOH* + e− | (15) |
M–OOH* + OH− → M + O2 + H2O + e− | (16) |
The difference between the Gibbs free binding energy of O* and OH* (ΔGO − ΔGOH) is used in the volcano-type correlation diagram of the OER reaction, which can evaluate the catalyst performance. Like the HER, the best catalyst displaying the minimum overpotential should bind intermediates on its surface neither too strongly nor too weakly. Cause the low potential, O* adsorption energy on the left side of the volcano (Fig. 1(c)) strong enough, the speed determination step is usually O* to OOH*. Inversely, on the right side of the volcano, the adsorption energy of O* is weaker, which resulted difficult formation of O* intermediates and this formation step is generally defined as decisive step. Among these major classes of metal oxide materials, all surfaces obey an unsatisfactory linear relationship between OOH* and OH* (ΔGOOH* = ΔGOH* + 3.2 ± 0.2 eV).
Another mechanism is the direct combination of 2 M–O intermediates to produce O2; after O2 release, the two empty metal centers are occupied by OH- to continue the reaction, which is known as the lattice oxygen participation mechanism (LOM). Unlike the traditional AEM mechanism system, LOM reactions occur at two adjacent metal sites rather than at a single metal site. Because *OOH is not an intermediate in the LOM catalytic cycle, limitation of the scaling relationship between *OH and *OOH can be broken. In the OER catalytic system, whether AEM or LOM mechanism, the study of structure–activity relationship is important, and continuous research is needed to develop more efficient and stable catalysis.
The typical LPE process comprises several important steps, as shown in Fig. 3(a),38 including immersion, insertion, exfoliation, and stabilization. By utilizing appropriate solvents and polymers, it is possible to prepare large quantities of monolayer or few-layer TMSes. For instance, commercially available TMSe powders such as MoSe2, TeSe2, and NbSe2 can be sonicated in various organic solvents with different surface tensions.39 After centrifuging the dispersion, the supernatant containing monolayer and few-layer TMSe nanosheets can be obtained. This method is insensitive to air and water, and has the potential for scalable production of exfoliated materials. Moreover, this procedure facilitates the formation of hybrid films with enhanced properties. Coleman reported a similar direct sonication method using a sodium cholate/water solution. However, direct sonication mechanically exfoliates layered structures, resulting in the breakage of nanosheets and reduction in lateral size. To address this issue, electrochemical exfoliation was developed to produce TMSe nanosheets with high yield and minimal damage. Furthermore, the bulk nanosheet and the insertion source during electrochemical exfoliation can serve as part of the electrode or electrolyte in a two-electrode system.
![]() | ||
Fig. 3 (a) Scheme of the whole LPE process containing immersion, insertion, exfoliation, and stabilization.38 Reproduced with permission from ref. 38. Copyright 2016, Wiley-VCH Verlag GmbH & Co. KGaA. (b) The electrochemical lithium interaction process.40 Reproduced with permission from ref. 40. Copyright 2017, Tsinghua University Press and Springer-Verlag Berlin Heidelberg. |
In electrochemical exfoliation, as shown in Fig. 3(b),40 the cation-intercalated substance can be inserted into the space between the layers of TMSe bulk materials, thereby increasing the interlayer distance and overcoming the van der Waals forces. In a typical electrochemical lithium intercalation process, first, the layered bulk material is initially mixed with acetylene black and a poly(vinylidene fluoride) (PVDF) binder dispersed in N-methylpyrrolidone (NMP). The resulting slurry is then uniformly coated onto a copper foil and dried. Subsequently, the bulk material-coated copper foil and lithium foil are used as the cathode and anode, respectively, with a polypropylene (PP) film serving as the separator. These components are assembled into a lithium-ion battery within an Ar-filled glove box, using a 1 M LiPF6 electrolyte in a mixture of ethylene carbonate (EC) and dimethyl carbonate (DMC) in a volume ratio of 1:
1.41
The performance of electrochemical exfoliation can be evaluated based on several parameters, including the amount of insertion, the voltage, and the discharging time. Firstly, the insertion amount. In the exfoliation process, the large sized of insertion can be formed under external voltage, which further increases the interlayer distance. For example, ultrathin ternary molybdenum sulfoselenide (MoSexS2−x) nanosheets can be prepared through cathodic electrochemical exfoliation in non-aqueous electrolytes. The exfoliated MoSexS2−x nanosheets exhibited high structural integrity, with lateral dimensions of up to ∼1.5 μm and an average thickness of approximately 3 nm.39 The cathodic electrochemical exfoliation of bulk MoSexS2−x was conducted in a two-electrode system, with bulk MoSexS2−x serving as the cathode, Pt foil as the anode, and TBAB solution in acetonitrile as the electrolyte. During exfoliation, the interaction of cations induced the deformation of bulk MoSexS2−x, resulting in the separation of ternary MoSexS2−x nanosheets from the parent MoSexS2−x and their dispersion in the TBAB acetonitrile solvent, exhibiting the Tyndall phenomenon. Secondly, the discharging voltage and time also influence the products of electrochemical exfoliation. In contrast to previously reported transition-metal sulfide crystals, which exhibit distinct discharge plateaus during electrochemical lithium intercalation, metal selenides show continuously descending discharge curves without a plateau. This makes it challenging to determine the cut-off voltage, which is the voltage at which the discharge stops, in order to achieve optimal Li insertion.42 Insufficient Li insertion can lead to ineffective exfoliation, while excessive Li insertion can result in the chemical decomposition of the crystals. In light of this, Zhang et al. successfully prepared few-layer NbSe2, WSe2, and Sb2Se3 nanosheets through electrochemical lithium intercalation at different cut-off voltages.41
In summary, for electrochemical exfoliation, the insertion amount, the voltage and the time of discharging can be carefully adjusted to achieve a high yield and maintain structural integrity of the layered TMSes.
Taking WSe2 as an example, the synthesis of WSe2 involves placing selenium powder upstream at approximately 20 cm from the center of the tube furnace, where WO3 is located. When the furnace is heated to 1000 °C with an appropriate ramping rate, the selenium vapor diffuses and reacts with the WO3 precursor, ultimately leading to the formation of WSe2 on the substrate.43 It is worth noting that, apart from WO3, ammonium salts containing tungsten can also serve as precursors. Du et al. have reported the successful preparation of WSe2 and W(SexS1−x)2 nanoflakes on carbon nanofibers using this approach.43 In their process, polyacrylonitrile nanofiber mats were placed on an alumina boat and heated to 280 °C for 6 h in the furnace tube to thermally decompose (NH4)6H2W12O40, resulting in uniformly dispersed WO3 nanoparticles. Subsequently, a boat loaded with Se and S powder was positioned 20 cm upstream from the pre-prepared WO3 nanoparticles at the center of the furnace. As the furnace temperature reached 1000 °C, the selenium vapor diffused and reduced the WO3 nanoparticles on the surface, forming volatile suboxides of WO3−x. These suboxides were rapidly selenized into the WSe2 product for lateral growth.
The properties and morphology of deposited TMSes can be effectively modified by precisely controlling the conditions during the CVD, including the choice of precursors and temperature. In the case of TMDs, their ultrathin sheet structure exposes prismatic edges and basal planes, while the termination atoms at the edges (metal or chalcogen) depend on the morphology, which is defined by the chemical potential of the growth environment. Therefore, by carefully controlling the synthesis and growth conditions during CVD, the edge termination of TMSes can be tuned, leading to alterations in their surface chemical properties. Traditional CVD methods always used transition metal oxides and got thin nanosheets; additionally, the morphology of TMSe products can be modified by using other precursors. For instance, yolk–shell structures of bimetallic MnCo selenide have been successfully prepared.44 In this process, Se powder and as-prepared MnCo glycolate powder were placed at the upstream and downstream sides of a tube furnace, respectively. The furnace was then heated to 450 °C and maintained to obtain the final product. As shown in Fig. 4(a), MoSe2 of different morphology can be fabricated through CVD under non-equilibrium conditions at different temperatures,45e.g., nanoplates were obtained at 880 °C, nanosheets at 940 °C, and microparticles at 980 °C. In the selenization process, the Se vapor can react with the target substrate without altering its original morphology. As shown in Fig. 4(c), this is exemplified by the preparation of pyrite-type beaded stream-like cobalt diselenide (CoSe2) nanoneedles directly on flexible titanium foils. This was achieved by treating a cobalt oxide (Co3O4) nanoneedle array template with selenium vapor.46 The property of TMSes is always connected with their morphology, and the choice of precursors and temperature have a significant effect on the practicle application of TMSes.
![]() | ||
Fig. 4 (a) Schematic diagram of the synthesis of MoSe2 structures by CVD at different temperatures; SEM images of the MoSe2 samples fabricated: nanoplates at 880 °C, nanosheets at 940 °C, and microparticles at 980 °C.45 Reproduced with permission from ref. 45. Copyright 2018, Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim. (b) Formation mechanism of macroporous CoSe2–CNT composite microspheres by the spray pyrolysis process and subsequent one-step post-treatment.49 Reproduced with permission from ref. 49. Copyright 2017, Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim. (c) Schematic depiction of the synthesis of CoSe2 nanoneedle arrays through treating a cobalt oxide (Co3O4) nanoneedle array template with selenium vapor. The experimental apparatus for step 2 (selenization) is shown as the inset.46 Reproduced with permission from ref. 46. Copyright 2013, The Royal Society of Chemistry. |
Heteroatom doping and heterojunction structures can be introduced into TMSes through the CVD method. For example, B-doped MoSe2 nanosheets can be synthesized using a two-zone furnace. Se powder was placed upstream of the B2O3 and MoO2 mixture, and subsequently heated to 550 °C and 920 °C.47 The NiSe2/Ni hybrid foam can be prepared by placing bare Ni foam and Se powder at the downstream and upstream of the furnace, ramping up to the anticipated temperature, and maintaining for 1 h under Ar flow.48 As shown in Fig. 4(b), Kim et al. synthesized CoSe2–CNT composite microspheres by placing the as-prepared CoO–CNT composite microspheres into a small alumina boat, which was loaded into a larger, covered alumina boat in a quartz reactor. The mixture was then heated to 300 °C for 6 h under a 10% H2–Ar reducing atmosphere.49 Selenium-enriched NiSe2 nanosheets can be converted from as-prepared ultrathin Ni(OH)2 nanosheets in a quartz socket. In the CVD process, Se powder and Ni(OH)2/CF were respectively placed upstream and downstream. After CVD growth, uniform and vertically textured NiSe nanosheets covering the entire CFs can be obtained.50 The introduction of heteroatoms or heterojunction structures in TMSes is significant for modifying their properties.
In the above studies, the Se and transition metal sources were placed separately. However, CVD selenization can also be achieved by uniformly mixing the reactants. For instance, the as-prepared Ni@nitrogen-doped graphene (NG) and Se powder can be annealed at 300 °C under a flow of argon for 3 h, thus obtaining a core–shell structure of NiSe2@NG.51 In addition to Se powder, didecyl selenide can also be used as a Se source. By heating the precursor mixture in a tubular furnace at 250 °C for 1 h under an N2 atmosphere, palladium selenide can be prepared.52
Several single-metal TMSes have been prepared by using the hydrothermal/solvothermal method, as shown in Fig. 5(a) and (b),53,54 such as MoSe2,53,55 FeSe2,56,57 and CoSe2,54,58 and in Fig. 3(a), such as NiSe2,59,60 CdSe,61,62 ZnSe,63 and CuSe.64 As shown in Fig. 5d, the MoSe2 nanosheets can be prepared by reacting Na2MoO4·2H2O with NaHSe at 220 °C for 12 h.55 Yin et al. have used excessive NaBH4 to regulate the electronic structure so that the MoSe2 framework was rearranged from 2H to 1T, and the MoSe2 nanosheets have the 1T phase for HER, as shown in Fig. 3(b).65 Fu et al. used NaHSe and ethanediamine to synthesize Ni0.85Se nanoparticles in the range of about 10–27 nm at 180 °C for 10 h.66 With Li2Se as the precursor, FeSe2 can be prepared by hydrothermal recrystallization.57 High-temperature conditions facilitated the formation of high-crystallinity CdSe hollow spheres with a wurtzite structure.64 The Co0.85Se nanosheets with a thickness of ∼30 nm can be obtained at 140 °C for 24 h, with Na2SeO3 as the Se source and water as solvent.58
![]() | ||
Fig. 5 (a) Schematic diagram of the synthesis of MoSe2/Ni0.85Se.53 Reproduced with permission from ref. 53. Copyright 2017, Elsevier Ltd. All rights reserved. (b) Solvothermal synthesis with CoSe2/DETA nanobelts as the substrate for the preparation of the MoS2/CoSe2 hybrid.54 (c) Schematic diagram of the morphology evolution of crystalline nickel selenide grown in situ on nickel foam.67 Reproduced with permission from ref. 67. Copyright 2013, The Royal Society of Chemistry. (d) Schematic illustration of phase- and disorder-controlled synthesis of MoSe2 nanosheets by the hydrothermal technique.55 Reproduced with permission from ref. 55. Copyright 2017, Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim. |
It is important to highlight that the quality of the product obtained from this method is significantly influenced by external conditions, including the reaction time, solvent, concentration of introduced additives, and so on. By adjusting the hydrothermal reaction time, nickel selenide arrays with various morphology can be achieved as shown in Fig. 5(c).67 Dai et al.68 utilized an organic precursor of selenium, namely selenium cyanoacetic acid sodium (NCSeCH2–COONa), where selenium is in the form of Se2−, without the use of a strong reducing agent or ethylene glycol.
Bimetallic selenides as well as ternary and multi-metallic selenides can be easily synthesized using hydrothermal/solvothermal methods. For instance, Zn-doped MoSe2 nanosheets can be prepared at 180 °C, which have been utilized as high-performance catalysts for the HER in acidic media.69 Mo–Fe selenide nanosheets with an ultrathin morphology (thickness about 20 nm) can be synthesized at 180 °C.70 Furthermore, Co–Fe–Se nanosheets with a thickness of about 2 nm can be obtained at 180 °C for 40 h by employing a mixed solution of EDTA and PVP.71
Various TMSe nanostructures such as heterostructure composites also can be formed by using the hydrothermal/solvothermal method. CoNiSe2 hetero-nanorods can be prepared at 160 °C.72 For example, CNT/Ni0.85Se–SnO2 networks can be prepared at 140 °C for 22 h.73 With hydrothermally synthesized NiMoO4 nanowires as the template, the MoSe2@Ni0.85Se nanowire network can be achieved at 160 °C for 24 h.53 Similarly, Zhang et al. used NiMoO4 nanowires to synthesize 3D hierarchical MoSe2/NiSe2 composite nanowires through a two-step hydrothermal approach.74 The hydrothermally prepared precursor was selenized at 220 °C for 12 h. The bare NF can be used as a substrate to grow 1D Ni–Se/NF nanostructures in different solvents, including ethanol, DMF and deionized water. With NaHSe solution as the Se source, the solvothermal process can be completed at 140 °C for 14 h.75 Li et al. grew the CoSe2/DETA nanobelts on Ti mesh at 180 °C for 16 h. The nanobelts exhibited widths of 30-300 nm and lengths of several micrometers.72 Yu et al. synthesized a CoSe2/DETA nanobelt substrate by adding Na2SeO3 and heating at 180 °C for 16 h.54 By introducing functional groups into MoSe2 decorated CBC nanofibers, Liu et al. synthesized a MoSe2/CBC composite by reacting Se powder with hydrazine hydrate, mixing it with Na2MoO4, and then heating at 180 °C for 12 h.76
In general, the hydrothermal/solvothermal method is a common method for preparing TMSes with different phases, components, and morphology, and the final product can be obtained on a large scale in a convenient and low-cost manner. Through rational design, various morphologies and structures can be achieved to enhance the catalytic performance.
In the electrochemical process, various substrates such as Si wafers, FTO (fluorine-tin-oxide) glass, metal foil, metal foam, and nanostructured templates are utilized as cathodes. The precursors are dissolved and electrodeposited onto the surface of the cathode. For instance, Co(C2H3O2)2, SeO2, and LiCl were added to a sealed electrochemical cell equipped with a graphite-rod counter electrode and an Ag/AgCl reference electrode. Amorphous cobalt selenide films were then electrodeposited onto a Ti substrate at a potential of −0.45 V.77 As shown in Fig. 6(a), Chia et al. prepared porous MoSex films on a glassy carbon electrode substrate by performing chronoamperometry at a certain potential for 10 min, using an aqueous electrolyte containing H2MoO4 and SeO2 precursors dissolved in NH4OH.78 The power supply during electrochemical deposition significantly impacts the structure of the resulting product. Copper selenide was synthesized through a single-step electrochemical deposition process using constant and pulsed potentials.79 Compared to potentiostatic deposition, the pulsed potential led to the formation of crystallized copper selenides. The electrochemical method has also been employed to prepare other TMSes, including MoSe2,80 WSe2,81 FeSe,82 NiSe2,83 CdSe,84 ZnSe,85,86 CuSe,79,87 MnSe,88 PbSe,89,90 and Ag2Se (Fig. 6(c)).91,92
![]() | ||
Fig. 6 (a) Illustration of the electrochemical fabrication of the inverse opal porous MoSe2 films.78 Reproduced with permission from ref. 78. Copyright 2018, American Chemical Society. (b) Schematics of MoSe2/Mo core–shell 3D hierarchical nanostructures fabricated by glancing angle deposition (GLAD) followed by the plasma-assisted selenization process.98 Reproduced with permission from ref. 98. Copyright 2016, Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim. (c) Schematic diagram of the CTL detection system.92 Reproduced with permission from ref. 92. Copyright 2010, Elsevier B.V. All rights reserved. (d) Schematic diagram of the experimental setup used for deposition of WSex films by SMPLD in a buffer gas.94 Reproduced with permission from ref. 94. Copyright 2013, Elsevier B.V. All rights reserved. |
Pulsed laser deposition has been utilized by Fominski and co-workers to fabricate MoSex (x = 1.5–2.4) thin films under various vacuum conditions.93 Similarly, WSex (x = 1.5–2.2) films can be prepared through shadow masked pulsed laser deposition (Fig. 6(d)).94 Fe-based TMSes superconducting films have been synthesized using molecular beam epitaxy, where ZnSe layers were grown on GaAs(001) under light illumination with the energy of approximately 1.8 eV.95 Additionally, other TMSes containing transition metals such as Cu and Zn have been obtained through pulsed laser deposition.96,97
The plasma-assisted selenization method has been reported to synthesize transition metal selenides. Diphase engineered 1T- and 2H-MoSe2/Mo catalysts can be prepared at a lower reaction temperature of 400 °C by using plasma to generate Se radicals.98Fig. 6(b) illustrates the selenization process, where Se pellets are positioned in the heater located at the top of the chamber and heated to 220 °C under a flow of N2/H2 mixture (N2:
H2 = 100
:
40, 2 Torr), and the plasma was turned on and maintained for 8 min once the substrate reached the target temperature.98
Other preparation methods have also been reported to synthesize TMSes, including cation exchange,35,99 photochemical method,100,101 sol–gel method,102,103 hot injection,104 one-pot synthesis,36,37 thermal evaporation,105,106 vacuum evaporation,107 spray pyrolysis,108,109 microwave assisted synthesis,110 sputtering, etc.111–113 These synthetic methods provide more means to acquire various morphology and enhanced performance, thereby maximizing the full potential of catalyst conditions. We have summarized these synthetic methods and related materials with morphology and phase, metal precursor source of selenides, method and reaction in Table 1. Nevertheless, there is still significant room to explore more facile methods and conditions to optimize TMSe-based materials.
No. | Material with morphology and phase | Metal precursor | Source of selenides | Method and reaction conditions | Ref. |
---|---|---|---|---|---|
1 | Few-layer 2H MoSe2, TeSe2, and NbSe2 | Commercial TMSe powders | — | Liquid phase exfoliation, centrifuging the dispersion | 39 |
2 | Few-layer-thick BN, NbSe2, WSe2, Sb2Se3, and Bi2Te3 | Layered bulk precursors | — | Electrochemical lithium intercalation process | 41 |
3 | Ultrathin ternary molybdenum sulfoselenide (MoSexS2−x) | Bulk MoSexS2−x | — | Cathodic electrochemical exfoliation | 39 |
4 | 2H WSe2 | WO3 precursor | Selenium powder | CVD | 43 |
5 | 2H WSe2 | (NH4)6H2W12O40 | Selenium vapor | CVD | 43 |
6 | Yolk–shell structures of bimetallic MnCo selenide | MnCo glycolate powder | Se powder | CVD | 44 |
7 | 2H MoSe2 nanoplates, MoSe2 nanosheets and MoSe2 microparticles | MoO2 precursor | Se powder | CVD process: nanoplates at 880 °C, nanosheets at 940 °C and microparticles at 980 °C | 45 |
8 | Pyrite-type beaded stream-like CoSe2 | Cobalt oxide (Co3O4) nanoneedle array | Se powder | CVD | 46 |
9 | B-doped MoSe2 nanosheets | B2O3 and MoO2 mixture | Se powder | CVD | 47 |
10 | NiSe2/Ni hybrid foam | Ni foam | Se powder | CVD | 48 |
11 | CoSe2–CNT composite microspheres | CoO–CNT composite microspheres | Se powder | CVD | 49 |
12 | Selenium-enriched NiSe2 nanosheets | Ni(OH)2/CF | Se powder | CVD | 50 |
13 | Core–shell structure of NiSe2@NG | Ni@nitrogen-doped graphene (NG) | Se powder | CVD | 51 |
14 | Palladium selenide | Palladium acetate | Didecyl selenide | CVD | 52 |
15 | MoSe2 nanosheets | Na2MoO4·2H2O | NaHSe | Hydrothermal | 55 |
16 | 1T phase MoSe2 nanosheets | Na2MoO4·2H2O | Se | Hydrothermal | 65 |
17 | 10–27 nm Ni0.85Se/NGO | Ni(NO3)2·6H2O | Se powder | Hydrothermal | 66 |
18 | Metastable stoichiometric FeSe2 | FeI2 | Li2Se | Hydrothermal | 57 |
19 | High-crystallinity CdSe hollow spheres with the wurtzite structure | Cd(NO3)2·4H2O | Se powder | Solvothermal | 64 |
20 | Co0.85Se nanosheets with a thickness of ∼30 nm | Na2SeO3 | Se source | Hydrothermal | 58 |
21 | Nickel selenide arrays | Nickel foam | Se powder | Solvothermal | 67 |
22 | MoSe2 hierarchical microspheres | (NH4)6Mo7O24·4H2O | NCSeCH2–COONa | Solvothermal | 68 |
23 | Zn-Doped MoSe2 nanosheets | Na2MoO4 | Se powder | Hydrothermal | 69 |
24 | Mo–Fe selenide nanosheets with an ultrathin morphology | FeCl3·6H2O, Na2MoO4 | Se powder | Hydrothermal | 70 |
25 | Co–Fe–Se nanosheets with a thickness of about 2 nm | CoCl2, FeCl3 | Selenium powder | Hydrothermal | 71 |
26 | CoNiSe2 hetero-nanorods | CoNi-precursor/NF | NaHSe solution | Hydrothermal | 72 |
27 | CNT/Ni0.85Se–SnO2 networks | CNT@NiSn(OH)6 nanocube precursor | Sodium selenite | Hydrothermal | 73 |
28 | MoSe2@Ni0.85Se nanowire | NiMoO4 nanowires | Se powder | Hydrothermal | 53 |
29 | 3D hierarchical MoSe2/NiSe2 | NiMoO4 nanowires | Se powder | Hydrothermal | 74 |
30 | 1D Ni-Se/NF nanostructures | Ni–Se/NF samples | NaHSe | Solvothermal | 75 |
31 | CoSe2/DETA nanobelts on Ti | Co(AC)2·H2O | SeO2 | Hydrothermal | 72 |
32 | CoSe2/DETA nanobelt substrate | Co(OAc)2·H2O | Na2SeO3 | Hydrothermal | 54 |
33 | MoSe2 decorated CBC nanofibers | Na2MoO4 | Se powder | Hydrothermal | 76 |
34 | Cobalt selenide films were electrodeposited on a Ti substrate | Co(C2H3O2)2 | SeO2 | Electrochemical | 77 |
35 | Porous MoSex films on a glassy carbon electrode substrate | H2MoO4 | SeO2 | Electrochemical | 78 |
36 | Copper selenide | Cu(NO3)2 | SeO2 | Electrochemical | 79 |
37 | MoSex (x = 1.5–2.4) thin films | MoSe2 target | — | Pulsed laser deposition | 93 |
38 | WSex (x = 1.5–2.2) films | Microcrystalline WSe2 | — | Shadow masked pulsed laser deposition | 94 |
39 | ZnSe layers on GaAs(001) | Zn fluxes | Se fluxes | Photo-molecular beam epitaxy | 95 |
40 | Diphase engineered 1T- and 2H-MoSe2/Mo | Mo film | Selenium pellets | Plasma-assisted selenization method | 98 |
Category | Catalyst | Electrolyte | HER | OER | Overall water splitting cell voltage (mV) at 10 mA cm−2 | Ref. | ||
---|---|---|---|---|---|---|---|---|
η (mV) (mA cm−2) (HER) | Tafel slope (mV dec−1) (HER) | η (mV) (mA cm−2) (OER) | Tafel slope (mV dec−1) (OER) | |||||
Single-metal selenides | c-CoSe2 | 1 M KOH | 190(10) | 85 | — | — | — | 124 |
c-CoSe2 | 0.5 M H2SO4 | 120(10) | 51 | 200(10) | 83 | — | 126 | |
CoSe and Co9Se8 | 1.0 M KOH | 268(100) | 61.4 | — | — | — | 13 | |
Thin CoSe2 | pH = 13 | — | — | 320(10) | 44 | — | 122 | |
Coral like CoSe2 | 1 M KOH | — | — | 295(10) | 40 | — | 123 | |
Flake-like Co7Se8 | 1 M KOH | 472(10) | 59.1 | 290(10) | 32.6 | 1.6 | 124 | |
d-WSe2/CFM | 0.5 M H2SO4 | 300(100) | 80 | — | — | — | 134 | |
1T-WSe2 | 0.5 M H2SO4 | 101 | 67 | 133 | ||||
Ni0.85Se | 1 M NaOH | 200(10) | 81 | 302(10) | — | — | 135 | |
Ni0.85Se | 0.5 M H2SO4 | 170(10) | 49.3 | — | — | — | 136 | |
Two-tiered NiSe | 1.0 M KOH | 177(10) | 58.2 | 290 | 77.1 | 1.69 | 142 | |
Textured NiSe2 | H2SO4 (pH ≈ 0.67) | 117(10) | 32 | — | — | — | 137 | |
NiSe–NiOx/NF | 1 M KOH | — | — | 274(16) | — | 1.74 | 138 | |
Ni0.75Se | 1 M KOH | 233(10) | 86 | — | — | 1.73 | 139 | |
NiSe2 | 1 M KOH | — | — | — | — | 1.43 | 140 | |
MoSe2 | 0.5 M H2SO4 | 300(10) | — | — | — | — | 127 | |
MoSe2 films with vertically aligned layers | 0.5 M H2SO4 | 159(10) | 61 | — | — | — | 128 | |
a-MoSe | pH ≈ 0 | 270(10) | 60 | — | — | — | 129 | |
pH ≈ 11 | 860(10) | 860 | — | — | — | |||
MoSex | 0.5 M H2SO4 | 570(30) | 118 | — | — | — | 78 | |
MoSe2/Mo core–shell nanoscrews | 0.5 M H2SO4 | 166 | 34.7 | — | — | — | 98 | |
Pd4Se | 0.5 M H2SO4 | 94 | 50 | — | — | — | 52 | |
PtSe2 | 0.5 M H2SO4 | 60 | 41 | — | — | — | 144 | |
InSe | pH = 1 | 549 | 126 | — | — | — | 149 | |
pH = 14 | 451 | 143 | ||||||
FeSe2 | 1 M KOH | 330 | 48.1 | — | — | — | 145 | |
2Fe–2Se | 1 M KOH | 245(10) | 1.73 | 146 | ||||
Heterogeneous structure | WSe2/CoSe2 | 0.5 M H2SO4 | 157 | 79 | 330 | 76 | — | 150 |
CoSe2|CoP | 1.0 M KOH | — | — | 240(10) | 46.6 | — | 151 | |
NiSe2–Ni2P/NF | 1.0 M KOH | 220(50) | 45 | 102(10) | 68 | 1.5 | 152 | |
MoSe2/NiSe2 | 0.5 M H2SO4 | 249(100) | 46.9 | — | — | — | 74 | |
MoSe2/Bi2Se3 | 0.5 M H2SO4 | 300(85) | 44 | — | — | — | 22 | |
(Ni,Co)Se2/CoSe2/NF | 1.0 M KOH | 65(10) | 140.2 | 255(10) | 148.6 | 1.56 | 205 | |
MOF-CoSe2@MoSe2 core–shell structure | 1.0 M KOH | 109.87(10) | 68.91 | 183.81(10) | 96.61 | 1.53 | 153 | |
Co(OH)2@CoSe nanorods (NRs) | 1.0 M KOH | 208(20) | 152 | 268(20) | 165 | — | 154 | |
MoSe2 HDH | 0.5 M H2SO4 | 191(10) | 72 | — | — | — | 40 | |
CoSe2–CNT | 0.5 M H2SO4 | 174(10) | 37.8 | — | — | — | 49 | |
Conductive substrate | CoSe2/CF | 1.0 M KOH | 95(10) | 52 | — | — | — | 155 |
Ni3Se2 on Au-coated Si | 0.3 M KOH | — | — | 290(10) | — | — | 93 | |
RGO/WSe2 | 0.5 M H2SO | 300(38.43) | 57.6 | — | — | — | 127 | |
Ni3Se2/NF-0.4 | 1.0 M KOH | 95(50) | 67 | — | — | 1.62 | 158 | |
NiSe-Ni0.85Se/CP | 1.0 M KOH | 101(10) | 74 | 1.53(10) | 98 | 1.53 | 156 | |
Ni0.85Se/rGO | 0.5 M H2SO4 | 104(10) | 50.7 | — | — | — | 66 | |
MoSe2 onto gold foil | 0.5 M H2SO4 | 107.2(10) | 31.8 | — | — | — | 80 | |
MoSe2/RGO | 0.5 M H2SO4 | 0.05(10) | 69 | — | — | — | 159 | |
MoSe2–rGO | 0.5 M H2SO4 | 0.21(10) | 57 | — | — | — | 160 | |
MoSe2/CNTs | 0.5 M H2SO4 | −0.07 V(10) | 58 | — | — | — | 161 | |
MoSe2–rGO–CNT | 0.5 M H2SO4 | 0.24(10) | 53 | — | — | — | 162 | |
N-doped RGO/MoSe2 | 0.5 M H2SO4 | 0.229(10) | 78.45 | — | — | — | 55 | |
CBC/MoSe2 | 0.5 M H2SO4 | 0.091 | 55 | — | — | — | 76 | |
NiSe2/NF | 1 M KOH | 104(10) | — | 279(20) | — | — | 157 | |
Defect engineering | Dual-native vacancies of Se and Mo in 2H-MoSe2 | 0.5 M H2SO4 | 126 mV(100) | 38 | — | — | — | 45 |
Defect modulation in WSe2 MLNSs | 0.5 M H2SO4 | 245(10) | 76 | — | — | — | 164 | |
1T phase-MoSe2 | 0.5 M H2SO4 | 152(10) | 52 | — | — | — | 116 | |
Ternary selenides | CoP2xSe 2(1 − x) | 1 M KOH | 98 | — | — | — | — | 166 |
0.5 M H2SO4 | 70 | — | ||||||
Meso-CoSSe | 0.5 M H2SO4 | 160(100) | 52 | — | — | — | 167 | |
W(SexS1−x)2 | 0.5 M H2SO4 | 110(10) | 59 | — | — | — | 168 | |
N-NiSe2/NF | 1.0 M KOH | 86(10) | 69 | — | — | — | 169 | |
Ni0.5Mo0.5Se | 0.5 M H2SO 4 | 197(10) | 107 | 340 | — | — | 170 | |
NiSe1.2S0.8 | 0.5 M H2SO 4 | 144(10) | 59 | — | — | — | 171 | |
N-doped MoSe2/TiC-C shell/core arrays | 0.5 M H2SO 4 | 137(100) | 32 | — | — | — | 206 | |
B-doped MoSe2 | 0.5 M H2SO 4 | 84(10) | 39 | — | — | — | 47 | |
MoSSe nanodots | 0.5 M H2SO 4 | 140(10) | 40 | — | — | — | 175 | |
Zn-doped MoSe2 | 0.5 M H2SO4 | 231(10) | 58 | — | — | — | 69 | |
MoSe2–Pt | 0.5 M H2SO4 | 250(10) | 85 | — | — | — | 176 | |
(Ni,Co)0.85Se nanoarrays | 1.0 M KOH | 216 | 77 | — | 177 | |||
NiCoSe2 nanosheet | 1.0 M KOH | 255.8(10) | 71 | 182 | ||||
(Ni,Co)0.85Se NSAs/NF | 1.0 M KOH | 169(10) | 115.6 | 287(20) | 86.7 | 1.65 | 183 | |
Co0.13Ni0.87Se2/Ti | 1.0 M KOH | 64(10) | 63 | 320(100) | 94 | 1.62 | 183 | |
CoNiSe2 (NRs) | 1.0 M KOH | 170(100) | 40 | 307(100) | 79 | 1.59 | 185 | |
Co-doped NiSe2 and Ni3Se4/C | 1.0 M KOH | 90(10) | 81 | 275(30) | 63 | 1.6 | 178 | |
Yolk-shelled Ni–Co–Se | 1.0 M KOH | 250 | 72 | 300 | 87 | — | 179 | |
CoxNi0.85−xSe | 0.5 M H2SO4 | 103(10) | 43 | — | — | — | 180 | |
CoNiSe2@CoNi-LDHs/NF | 1 M KOH | 106(10) | 74 | — | — | — | 72 | |
(Ni,Co)Se2-GA | 1 M KOH | 128(10) | 79 | — | — | — | 181 | |
(Ni,Fe)3Se4 | 1 M KOH | 225(10) | 41 | — | — | — | 189 | |
((Ni0.75Fe0.25)Se2)/CFC | 1 M KOH | 255(35) | 47.2 | — | — | — | 192 | |
FeNi2Se4–NrGO | 1 M KOH | 170(10) | 62.1 | — | — | — | 193 | |
(CoMn)Se2 | 1 M KOH | 270(10) | 39 | — | — | — | 195 | |
Multi-metal selenides | Fe0.37Ni0.17Co0.36Se | 1 M KOH | 179(10) | 115.3 | — | — | — | 65 |
NixFe1−xSe2-DO | 1 M KOH | 195(10) | 28 | — | — | — | 186 | |
Ni0.75Fe0.25Se2 | 1.0 M KOH | 255 | 56 | — | 187 | |||
Fe-doped NiSe | 1.0 M KOH | 163(10) | 71.4 | 231(50) | 43 | 1.585 | 190 | |
NiFeSe@NiSe|O@CC | 1.0 M KOH | 62(10) | 48.9 | 270(10) | 63.2 | 1.56 | 191 | |
Ni-FeSe | 1 M KOH | 197(10) | 56 | — | 188 | |||
(Fe–Co)Se2 | 1 M KOH | 251(10) | 47.6 | 207 | ||||
Co1−xFexSe2 | 1 M KOH | 217 | 41 | — | 194 | |||
MoCoSex@NC | 0.5 M H2SO4 | 60(10) | 67 | — | — | — | 196 | |
(Co0.21Ni0.25Cu0.54)3Se2 | 1.0 M KOH | 272(10) | 53.3 | — | 201 | |||
(Fe0.48Co0.38Cu0.14)Se | 1.0 M KOH | 256(10) | 40.8 | — | — | — | 200 | |
ZnNi0.5Co0.5Se2/Cu1.8Se | 1.0 M KOH | — | — | 370(325) | 72 | — | 202 | |
Ni–Co–Fe–Se@NiCo-LDH | 1.0 M KOH | 113(10) | 44.87 | 286(100) | 78.9 | — | 203 | |
Fe0.09Co0.13–NiSe2 | 1.0 M KOH | 251(10) | 92(10) | 17 | ||||
NiFeCoSex/CFC | 1.0 M KOH | 150(10) | 85 | — | — | — | 199 |
Among various cobalt selenides for HER electrocatalysis, the most representative one is cobalt diselenide (CoSe2), which exists in three distinct crystal structures: orthorhombic marcasite-type (o-CoSe2), cubic pyrite-type (c-CoSe2, electronic configuration t2g6eg1), and polymorphic CoSe2 (p-CoSe2) with mixed orthorhombic and cubic phases. The electrocatalytic activity of c-CoSe2 for HER under alkaline conditions surpasses that of other CoSe2 materials.124 As shown in Fig. 7,124 the performance of o-CoSe2 and c-CoSe2 was compared through both experimental and DFT calculations. The water adsorption energy of the c-CoSe2 catalyst (−0.163 eV) was significantly lower than that of the o-CoSe2 catalyst (−0.106 eV), indicating easier water absorption on the surface of c-CoSe2 in alkaline media. Theoretical investigations proved the easier water-adsorption process on cobalt cations and faster transformation of Hads into H2 on the Se active sites. Additionally, the electrical conductivity of c-CoSe2 was also better. Consequently, the polarization curve of c-CoSe2 achieved a low overpotential of 190 mV at a current density of 10 mA cm−2.
![]() | ||
Fig. 7 (a) The XRD patterns of c-CoSe2/CC and o-CoSe2/CC and (b) Raman spectra of c-CoSe2 and o-CoSe2 products. (c) and (d) The SEM images of c-CoSe2/CC and o-CoSe2/CC 3D electrodes. Inset: Corresponding high-resolution SEM images. (e) and (f) The HRTEM images of c-CoSe2 and o-CoSe2 products. Inset: Corresponding selected area electron diffraction (SAED) patterns. (g) Calculated HER free-energy change and (h) water adsorption energy for the o-CoSe2 and c-CoSe2 products. (i) IR-corrected polarization curves and (j) corresponding Tafel slopes of blank CC, Co(OH)F/CC, c-CoSe2/CC, and o-CoSe2/CC.124 Reproduced with permission from ref. 124. Copyright 2016, Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim. |
As bi-functional catalysis for water splitting, cobalt selenides are also considered as high performance catalysts for both HER and OER. Masud et al.125 have prepared Co7Se8 nanostructures with a flake-like morphology which show bifunctional catalytic activity for OER and HER in alkaline medium with long-term durability (>12 h) and high faradaic efficiency (99.62%). Co7Se8 exhibited an excellent OER activity in 1.0 M KOH with a small overpotential of 290 mV at a current density of 10 mA cm−2. Moreover, this catalyst was found to be effective for hydrogen evolution reaction in the same media. It can effectively split water at a cell voltage of 1.6 V under ambient conditions. The catalytic performance can be further enhanced by modifying the morphology. To synthesize c-CoSe2 nanoparticles embedded in carbon cages, Co-based metal–organic frameworks (MOFs) were utilized as precursors.126 By controlling the annealing conditions, the phase transformation of CoSe2 from orthorhombic to cubic phase was achieved. Due to its metallic nature, cubic phase CoSe2 exhibits significantly improved electrocatalytic activities for both HER (η = 43 mV, Tafel slope = 51 mV dec−1) and oxygen evolution reaction (OER) (η = 200 mV, Tafel slope = 83 mV dec−1). The porous structure of the MOF cages not only exposes the active sites of incorporated CoSe2, but also prevents the catalysts from agglomerating during the catalysis. Manipulating the charge state of Co species and controlling the mass ratio of Co and Se also can enhance the catalytic performance. Zhao et al. synthesized 3D free-standing cobalt selenide electrodes containing CoSe and Co9Se8 phases through one-step calcination of Co foil with Se powder in a vacuum-sealed ampoule.13 The electrocatalytic behaviors of CoSe2 can be significantly affected by the change of Co charge states. Synchrotron-based mechanism study has demonstrated that a high Co charge state is beneficial for OER, while a low Co charge state promotes the HER activity. The 3D network structure of cobalt selenide electrodes facilitates efficient charge transfer, leading to enhanced and stable electrocatalytic performance.
The morphology engineering has been able to increase the active sites. Ravikumar et al. synthesized MoSe2 nanoflowers, which exhibited a higher number of exposed chalcogen sites and superior conductivity, contributing to their high performance in acidic HER.127 As shown in Fig. 9(a)–(c), MoSe2 films with vertically aligned layers were prepared through CVD, resulting in active sites located at the edges of MoSe2.128 Based on an electrochemical corrosion process, the amorphous molybdenum selenide can be activated under H2-evolving conditions, and produce a steady overpotential of 270 mV at 10 mA cm−2.129 Chia et al. fabricated inverse opal porous MoSex films through electrosynthesis, where 2 ≤ x ≤ 3. In this process, MoSex was electro-deposited via a solid template-assisted method.78 Molybdic acid and selenium dioxide were simultaneously reduced as the respective metal and chalcogen precursors in an aqueous electrolyte. After removing the template, the electrosynthesized porous MoSex films contained pores with diameters of 0.1, 0.3, 0.6, or 1.0 μm, depending on the size of the template. The porous MoSex films with a pore size of 0.1 μm exhibited the lowest HER overpotential of 0.57 V at −30 mA cm−2 and a Tafel slope of 118 mV dec−1. Phase-engineered 1T- and 2H-MoSe2/Mo 3D-hierarchical nanostructures with controlled shapes were prepared using a low-temperature plasma-assisted selenization process.98 The metallic 1T-MoSe2/Mo core–shell 3D-hierarchical nanostructures were utilized for HER measurements and exhibited the highest catalytic activities. The increased density of exposed edges and the metallic phase of the MoSe2 shell contributed to the high performance of the 1T-MoSe2/Mo core–shell 3D-hierarchical nanostructures. The design and assembly of multi-dimensional structures can serve as a promising approach for the development of cost-effective and efficient TMSe catalysts for HER applications.
MoSe2 and WSe2 as typical 2D-layered transition metal dichalcogenides and promising electrocatalysts to replace noble metals have been widely studied. Currently, because of the relatively wide band gap and large basal plane of MoSe2 and WSe2, they are mostly designed and applied for HER. For OER, the catalysts should possess high conductivity to meet the complicated kinetics. The 1T phase has metallic behavior, but its stability is poor. It is very important to find a suitable method to improve its stability for water splitting. Most MoSe2 (WSe2)-based electrocatalysts show good HER activity. The HER/OER performance in alkaline media is still not excellent.
![]() | ||
Fig. 8 (a) Schematic illustration of the synthesis of Ni0.85Se on a graphite substrate. (b) XRD patterns of the as-synthesized Ni0.85Se/GS sample calcined at 400 °C. (c) A typical SEM image of Ni0.85Se grown on GS. (d) A low magnification TEM image of the Ni0.85Se catalyst. (e) A high resolution TEM image of the Ni0.85Se catalyst and the corresponding fast Fourier transform (FFT) pattern (inset). Electrocatalytic performance of Ni0.85Se samples calcined at 400 °C in 1 M NaOH. (f) LSV curves of Ni0.85Se/GS and graphite for HER. Inset: Tafel plot of Ni0.85Se/GS derived from the LSV curve for HER. (g) LSV curves of Ni0.85Se/GS and graphite for OER. Inset: Tafel plot of Ni0.85Se/GS derived from the LSV curve for OER. (h) and (i) 48 h long-term stability test of Ni0.85Se under constant current densities of 10 mA cm−2 for HER and OER. (j) LSV curve of water electrolysis for Ni0.85Se/GSNi0.85Se/GS and RuO2Pt/C. (k) Chronopotentiometric curve of Ni0.85Se/GS. Inset: A photograph of the system showing hydrogen and oxygen generated during the durability test.135 Reproduced with permission from ref. 135. Copyright 2016, Hydrogen Energy Publications LLC. Published by Elsevier Ltd. All rights reserved. |
Nickel sulfide as a promising OER electrocatalyst, the long bond length can weaken the Ni–Ni bond, promote surface oxidation, and improve the catalytic performance. Therefore, nickel selenide was thought to exhibit high performance towards OER reactions. In view of the easily oxidized nature of selenides under strong alkaline conditions, the possible electrochemical surface re-construction of the selenides during electrocatalysis is often neglected. Gao et al.138 conducted a thorough and meticulous study on the NiSe oxygen evolution nanoelectrocatalyst during the OER process. In parallel with the OER process, Se2− in NiSe loses electrons to form the soluble selenoxide species, whereas Ni2+ evolves into NiOx. Once the NiOx layer formed on the NiSe surface reaches a certain thickness, it will protect the inner NiSe from further oxidation. These results further imply that the electrocatalytic activity of the so-called NiSe catalysts should originate from the amorphous NiOx shell, with the NiSe core having some auxiliary effects on the catalytic activity. In addition, the NiSe–NiOx/NF material can serve as an efficient electrocatalyst for HER. They tested the water-splitting activity in 1 M KOH, and the current density was achieved at 1.74 V (20 mA cm−2).
Nickel selenides with non-stoichiometry such as Ni0.85Se have unsaturated atoms, which is of great significance to explore the electrocatalytic performance in water splitting. To understand the effects of the Se/Ni ratio of non-stoichiometric nickel selenides on the electrocatalytic properties (Fig. 9(d)–(g)), Zheng et al. developed a novel hot-injection process at ambient pressure to control the phase and composition of a series of NixSe by simply adjusting the added molar ratio of the nickel resource.139 Among the synthesized NixSe series, the cubic nanocrystalline Ni0.5Se exhibits superior OER activity comparable to RuO2 (330 mV, 10 mA cm−2), which may benefit from the pyrite-type crystal structure and Se enrichment in Ni0.5Se. Ni0.75Se is proved to be most efficient for HER, with an overpotential of 233 mV at an HER current density of 10 mA cm−2; the remarkable catalytic activity of Ni0.75Se benefits from the good electronic conductivity and enhanced electrochemically active area. Ni0.75Se as the cathode and Ni0.5Se as the anode require a cell voltage of 1.73 V (10 mA cm−2) for overall water splitting. The analysis demonstrates that the electronic conductivity of Se/Ni and Se content in the NixSe compounds account for the electrocatalytic activity. In addition to non-stoichiometric nickel selenide compounds, stoichiometric NiSe2 has a pyrite structure with a zero bandgap. Swesi et al. prepared NiSe2 films through electrodeposition for use as an efficient bifunctional electrocatalyst for overall water splitting.140 The final product could effectively split water at 1.43 V and achieve an electrolysis energy efficiency of 83% producing current density as high as 100 mA cm−2. The high performance is attributed to lowering of the oxidation potential of Ni2+ to Ni3+ which in turn is a consequence of changing the oxide lattice to the more covalent selenide lattice, as well as the preferential growth direction of the film which exposes the Ni-rich surface as the terminating lattice plane. The electrocatalytic performance can be further improved by increasing the number of active reaction sites. To provide more active sites, the surface roughness of NiSe2 was enhanced using an acetic acid-assisted method. Zhang et al. successfully prepared porous NiSe2 nanowrinkles anchored on nickel foam (NF), which exhibited efficient electrochemical activity for both HER and OER.141 Wu et al. developed ultrathin hierarchical nanosheets (≈0.96 nm) by a self-limiting tunable acid etching and topotactic selenization process.142 Through rationally controlled hydrolyzation with the self-regulated pH value, Ni foam was etched into ultrathin layered Ni(OH)2. This ultrathin structure prevented excessive lattice expansion in bulk materials during selenization. Consequently, Ni(OH)2 was artificially transformed to ultrathin NiSe (≈1.25 nm) through a simple and nondestructive topotactic phase engineering approach. Thus, highly efficient and stable OER and HER electrodes with low Tafel slopes and onset potentials can be achieved under alkaline conditions.
![]() | ||
Fig. 9 (a) Schematic illustration of the preparation of perpendicularly oriented TMD nanosheets/GN hybrids. (b) and (c) SEM images of MoSe2 nanosheets grown on GN with a graphite disc as the substrate.128 Reproduced with permission from ref. 128. Copyright 2016, Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim. Crystal structures of (d) NiSe, (e) Ni0.85Se, (f) Ni0.75Se, and (g) Ni0.5Se nanocrystals.139 Reproduced with permission from ref. 139. Copyright 2018, American Chemical Society. (h) Schematic of the HER measurement setup. (i) Polarization curves for the HER obtained on the basal plane, the edge and the whole flake. Inset: Optical image of a typical PMMA masked electrochemical microcell. (j) Optical images of CVT-grown PtSe2 flakes with one to five and ∼20 layers, insets: corresponding AFM images of the PtSe2 flakes with height profiles. (k) Polarization curves and Tafel plots for the HER obtained on 1–5 L and ∼20 L PtSe2 flakes and a commercial Pt/C catalyst.144 Reproduced with permission from ref. 144. Copyright 2019, Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim. |
![]() | ||
Fig. 10 (a) The exfoliation energies Eexf as a function of interlayer spacing D. (b) The Gibbs free energy ΔGH diagrams for the pristine 3d-TMSe. (c) The ΔGH summary of 3d-TMSe, 3d-TMSe/V1 and 3d-TMSe/V2, wherein V1 and V2 stand for the single cationic-vacancy and dual cationic-vacancy pair. The red dashed line indicates the screening criterion of the outstanding performance for HER catalysis. (d) The partial density of states (PDOS) of CuSe, CuSe/V1 and CuSe/V2. (e) The vacancy formation energies Ef for 3d-TMSe/V1 and 3d-TMSe/V2.143 Reproduced with permission from ref. 143. Copyright 2020, Elsevier B.V. All rights reserved. |
Pt and Pd are common elements used in energy conversion. They can also be synthesized as TMSe-based electrocatalysts for water splitting. For example, various palladium selenides (Pd4Se, Pd17Se15, Pd7Se4) can be prepared through thermal decomposition using the same palladium and selenium precursors with different mole ratios. Among them, Pd4Se exhibited the best HER catalytic activity in acidic media, with an overpotential of 94 mV to drive a current density of 10 mA cm−2, an exchange current density of 2.3 × 10−4 A·cm−2, an apparent Tafel slope of 50 mV dec−1, and a turnover frequency (TOF) value of 0.47 s−1.52 The combination of intrinsic lower charge transfer, larger electrochemical surface area, and moderate proton binding energy contributed to the high performance of palladium selenides. Furthermore, the Pd4Se phase with tetragonal symmetry which possessed a higher d-band DOS had a smaller positive charge on Pd compared to the other two cases. Therefore, the oxidation state of Pd in Pd4Se is smaller than that of the other two phases. This will help in the effective adsorption of protons on Pd4Se as compared to Pd7Se4 and Pd17Se15. PtSe2 demonstrates high catalytic activity for HER and notable layer-dependent electrical properties.144 However, it is challenging to accurately characterize the layer structure using traditional electrochemical methods. Jiao et al. synthesized single crystalline 2D PtSe2, as shown in Fig. 9(h)–(k), with a controlled number of layers and examined the HER catalytic activity of individual flakes using micro electrochemical cells. The number of layers influenced the number of active sites, electronic structures, and electrical properties of the 2D PtSe2 flakes, thereby affecting their HER performance. This study precisely investigated the relationship between the layer number of nanosheets and their catalytic performance using micro electrochemical cells, significantly enhancing our understanding of the structure–activity correlations for 2D nanosheets.
In addition to Pt and Pd, other non-noble metals like Cu can also constitute TMSe-based electrocatalysts. DFT calculations identified CuSe as a promising candidate toward HER.143 The pristine CuSe monolayer, without any treatment, exhibits relatively good activity, as indicated by the small thermodynamic barrier of 0.3 eV. In contrast, the Se site on the pristine basal plane of the remaining 3d-TMSes (including Ti, V, Cr, Mn, Fe, Co, Ni, and Zn) is chemically inactive. FeSe2 has a relatively small Tafel slope among Fe, Co and Ni based dichalcogenides and offers an opportunity to improve the performance of nonmetallic catalysts by adjusting the structure. Generally, TMSes share a similar structure with the corresponding sulfides but have higher electrical conductivity, resulting in superior charge transfer in selenides such as Fe- and Co-based selenides. Gao et al.145 have synthesized 2D FeSe2 nanoplatelets via a hydrothermal reduction route. The FeSe2 nanoplatelets exhibit highly enhanced OER catalytic activity compared with commercial RuO2 (e.g., overpotential: 2.2 times higher than that of commercial RuO2 at 500 mV; Tafel slope: 48.1 mV dec−1; steady-state current densities remain constant after 70 h). The excellent catalytic activity can be ascribed to the 2D nanostructure, which could facilitate improvement of kinetics of water oxidation, and the high density of exposed active sites on the (210) crystal surface. Density functional theory calculations provide detailed insight into the energy level and water adsorption at the surface of FeSe2 nanoplatelets.
Therefore, experimental and theoretical studies prove that FeSe2 with intrinsic semiconductor properties and tunable electronic structures effectively reduces the OER overpotential. Iron selenides can serve as catalysts in HER, OER, and overall water splitting. Panda et al.146 have investigated whether FeSe2 can be fine-tuned to achieve high performance not only for bifunctional OER and HER but also for overall water-splitting, through the control of size, structure, morphology, and electronic properties. They have synthesized a bioinspired molecular 2Fe–2Se cluster core supported by a β-diketiminato ligand via a versatile thermolytic approach. The as-synthesized FeSe2 was electrophoretically deposited on nickel foam and possessed an overpotential of 245 mV at a current density of 10 mA cm−2, representing outstanding catalytic activity and stability due to the formation of Fe(OH)2/FeOOH active sites at the surface of FeSe2. Since the medium bonding with intermediates and products in HER results in good catalytic activity, Se–H bond strength is lower than that of S–H, so that it can act as a base to trap protons and facilitate deprotonation to release hydrogen in HER. Moreover, an overall water splitting setup was fabricated using a two-electrode cell which showed a small cell voltage and high durability. Such bioinspired molecular 2Fe–2Se cluster was creatively synthesized and as foundation for the application of iron selenide in overall water splitting. The MnSe electrode prepared on nickel foam showed an overpotential of 311 mV at 10 mA cm−2 in an alkaline electrolyte. With the incorporation of iron into MnSe, the MnFeSe electrode demonstrated significantly enhanced OER activity and improved reaction kinetics.147 To gain insight into the high OER/PEC activities of the Mn-based catalysts, especially the intermediate phase of MnSe, density functional theory (DFT) calculations were performed on six surface models of MnSe, Fe–MnSe, MnOOH, Se, Fe–MnOOH, MnO2, and Se, Fe–MnO2. It was found that both the rate-determining steps on MnSe and Fe–MnSe are O* + OH → OOH* + e−. Additionally, it was inferred that the Fe dopant can stabilize the O species, accelerating the MnSe oxidation and Mn (oxy)hydroxide formation of MnOOH and δ-MnO2 after the OER. GaSe and GeS also possess unique electrochemical properties.148 GeS is isoelectronic to black phosphorus with the same structure. GaSe is a layered material composed of GaSe sheets bonded in the sequence of Se–Ga–Ga–Se. It was found that the encompassing surface oxide layers on GaSe and GeS greatly influenced their electrochemical properties, especially their electrocatalytic capabilities towards HER. These findings provide new insights into the electrochemical properties of these IIIA–VIA and IVA–VIA layered structures. InSe, with small flakes, exhibits predominant edge effects. Elisa Petroni et al. produced few-layered InSe flakes through liquid-phase exfoliation of β-InSe single crystals in 2-propanol with a low boiling point (82.6 °C). These flakes have maximum lateral sizes ranging from 30 nm to a few micrometers and thicknesses ranging from 1 to 20 nm. The HER performance of InSe was tested in hybrid single-walled carbon nanotubes/InSe heterostructures. InSe flakes with smaller surface areas and thicknesses exhibited the lowest reported values of η10 among the MX compounds (η10 = 549 mV at pH = 1 and η10 = 451 mV at pH = 14).149
CoSe2 nanoparticles were effectively dispersed in WSe2 nanosheets through a hot-injection colloidal synthesis process for refining their electronic functionalities, as shown in Fig. 11.150 The resulting WSe2/CoSe2 heterostructured nanohybrids exhibited enhanced electrocatalytic performance for both HER and basic OER, with lower overpotential values (η10), and Tafel slope values of 157 mV and 79 mV dec−1 for HER, and 330 mV and 76 mV dec−1 for OER, respectively. The improved electrocatalytic performance can be attributed to enhanced conductivity and increased surface area. Furthermore, the hybrid catalyst of CoSe2|CoP was synthesized and applied for OER.151 It was found that the true OER-active species in alkaline medium were oxyhydroxides and hydroxides of CoSe2 and CoP, rather than selenides and phosphides. Selenides got oxidized to oxyhydroxides/hydroxides amid of OER, whereas phosphides took an extended period to fully oxidize. As a result, the derived oxides from the hybrid catalyst exhibited a relatively low overpotential (Fig. 12).
![]() | ||
Fig. 11 (a) Schematic representation showing the formation of WSe2/CoSe2 heterostructured nanohybrids. (b)–(d) HRTEM images, (e) SAED pattern, (f) HAADF-STEM image, and (g)–(i) related EDX mapping images for WSe2/CoSe2 nanohybrids. (j) HER polarization profiles and (k) the corresponding Tafel plots of the WSe2/CoSe2 nanohybrids. (l) OER polarization curves and (m) the corresponding Tafel plots of the WSe2/CoSe2 nanohybrids.150 Reproduced with permission from ref. 150. Copyright 2021, American Chemical Society. |
![]() | ||
Fig. 12 (a) Schematic illustration of the preparation of CoSe2|CoP by solvothermal and phosphidation processes. (b) XRD data of CoSe2|CoP. (c) iR-corrected polarization curves of CoSe2–DO, CoP–DO, and CoSe2|CoP–DO electrodes, along with RuO2 for comparison. (d) Overpotential (η) required for j = 10 and 100 mA cm−2. (e) Tafel plots derived from the corresponding polarization curves of CoSe2–DO, CoP–DO, and CoSe2|CoP–DO. (f) Overpotential required at 10 mA cm−2 (η10) and Tafel slope comparison of the catalysts in this work with other recently reported high-performance OER electrocatalysts.151 Reproduced with permission from ref. 151. Copyright 2018, Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim. |
Nanowrinkles of NiSe2 and Ni2P on Ni foam (NiSe2–Ni2P/NF) were synthesized through full selenization and partial phosphorization.152 In alkaline solution, NiSe2–Ni2P/NF exhibited favorable catalytic performances for both OER and HER. The heterostructure of NiSe2 and Ni2P demonstrated a synergistic effect on the water adsorption process and catalytic kinetics enhancement. Moreover, this catalyst functioned as a bifunctional electrode for overall water splitting, requiring a cell voltage of only 1.50 V to achieve a current density of 10 mA cm−2. In a different approach, MoS2/CoSe2 structures were integrated to achieve a synergistic effect between MoS2 and CoSe2. The catalyst exhibited an increased number of catalytic sites and reduced free energy, resulting in a Tafel slope of 36 mV dec−1.54
3D hierarchical MoSe2/NiSe2 composite nanowires (NWs) were prepared on a carbon fiber paper (CFP) skeleton.74 The nanosheets of MoSe2 and NiSe2 NWs were uniformly and tightly distributed on CFP. Consequently, the formed 3D hierarchical MoSe2/NiSe2 delivered a current density of 100 mA cm−2 with an overpotential of 249 mV and a Tafel slope of 46.9 mV dec−1. In the electrocatalytic process, the 3D hierarchical structure could effectively suppress the aggregation or re-stacking of nanosheets, exposing more active sites to participate in HER.
Additionally, the highly conductive NiSe2 dispersed in the nanosheets facilitated electron transfer from the electrode to the active edges of MoSe2, further enhancing the activity. Hetero nanostructures can also be prepared by evenly loading small ultrathin hierarchical MoSe2 nanosheets on the whole surfaces of single-layer Bi2Se3 hexagonal nanoplates. The MoSe2/Bi2Se3 hybrids showed an overpotential (η) of 300 mV at a current density of 85 mA cm−2, and a Tafel slope of 44 mV dec−1.22 Another heterostructure electrocatalyst of (Ni,Co)Se2/CoSe2/NF was fabricated on NF and applied as a superior bifunctional electrode for overall water splitting. The as-synthesized (Ni,Co)Se2/CoSe2/NF electrocatalyst exhibited excellent electrochemical performance in alkaline solutions with HER and OER overpotentials of 65 mV and 255 mV, respectively, at a current density of 10 mA cm−2. The alkaline electrolyzer for overall water splitting required a cell voltage of only 1.56 V to drive 10 mA cm−2 current. The hybrid core–shell structure was another kind of heterogeneous structure; MoSe2 2D nanosheets were uniformly grown on a conductive Co-metal–organic framework (Co-MOF), resulting in cobalt diselenide laminated with molybdenum diselenide.153 Such core–shell structure exhibited low overpotentials and a low voltage for overall water splitting. In the developed nanostructure, Co-MOF was converted into CoSe2, providing conductive unidirectional pathways between the MoSe2 nanosheets and carbon cloth (CC). The decoration of the MoSe2 nanoflakes on the CoSe2 nanowalls improved the resultant electroactive sites for the electrochemical process, and the porous 2D nanosheets of MoSe2 allowed intimate contact with the electrolyte, and so rich reaction sites can be used more efficiently in core–shell structures which minimizes the ion diffusion length.
Hydroxides and TMSes also can constitute a heterogeneous structure to further enhance the activity. For instance, Co(OH)2@CoSe nanorods were evaluated as electrocatalysts for both HER (overpotential of 208 mV at 20 mA cm−2) and OER (268 mV at 20 mA cm−2) at high current density in a 1 M KOH solution.154 The nanorods exhibited superior electrocatalytic activity, attributed to the chemical coupling of Co–OH active sites between Co(OH)2 and CoSe, which regulated hydrogen adsorption and desorption energy, and facilitated fast electron transfer. Additionally, the hetero-dimensional layered MoSe2 nanodots (NDs) anchored on few-layer MoSe2 nanosheets (NSs) (MoSe2 HDH) via a one-pot hydrothermal route were designed and prepared as efficient electrocatalysts for HER.40 MoSe2 NDs with edge-abundant features and the defect-rich structure could introduce more active sites for HER. Random stacking of the flake-like MoSe2 NSs on the surface of the supporting electrode can facilitate electron transport efficiency.
Damien et al. have developed a facile approach for depositing monodisperse single/few layers of MoSe2 onto gold foil through electrochemical exfoliation. The resulting catalyst exhibited exceptional electro-catalytic activity towards the HER with a Tafel slope of 31.8mV dec−1, i.e., a Pt-like activity for HER.80 Overall, the research highlights the potential of TMSe catalysts for HER and the importance of a conductive substrate in improving performance. The substrate choice also has a significant effect on the OER catalytic performance. Ni3Se2 which was electrodeposited on different conducting substrates, including Au-coated glass, Au-coated Si, glassy carbon (GC),93 ITO-coated glass, and Ni foam, has been synthesized and investigated for the OER activity. It was observed that it can generate 10 mA cm−2 at an overpotential as low as 290 mV (upon annealing or using Au-coated Si as the growth substrate), and verified that the OER activity was influenced by electrodeposition parameters including the deposition time and pH of the electrochemical bath, annealing, and the nature of the substrate. In addition to metallic substrates, carbon-based, Ni foam and bio-substrate materials have been utilized to enhance conductivity and prevent the aggregation of different TMSes.
For CoSe2, Dong et al. have reported a facile method for synthesizing CoSe2 nanoparticles anchored on flexible carbon fiber (CF) electrodes via pyrolysis and selenization of ZIF-67 directly grown on CFs (CoSe2/CF).155 This CoSe2/CF composite also exhibited excellent catalytic activity (95 mV at 10 mA cm−2, Tafel slope of 52 mV dec−1) and stability in alkaline media. Specifically, Kang et al. have developed an orthorhombic CoSe2–CNT composite with an optimized morphological structure through spray pyrolysis and selenization.49 This composite exhibits excellent catalytic activity for HER in an acidic medium. However, the bare CoSe2 powders show moderate HER activity with an overpotential of 226 mV at 10 mA cm−2, and Tafel slopes of 58.9 mV dec−1. The high performance of the CoSe2–CNT composite is attributed to its macroporous structure resulting from the CNT backbone, which has effectively improved the contact area of CoSe2 active sites with the liquid medium, enhanced the H2 removal rate, and increased electrical conductivity.
For WSe2, to further increase the exposed active sites and electrical conductivity, Zhang et al. constructed an RGO/WSe2 hybrid network and anchored it on graphene nanosheets.127 Compared with bare WSe2 nanoflowers with relatively poor HER performance, the RGO/WSe2 hybrid demonstrated highly effective and stable electrocatalytic performance.
For NiSe, as shown in Fig. 13(a)–(h), Fu et al. have reported hetero-phase junction catalysts containing a hexagonal phase on carbon paper (CP), namely, NiSe/CP, Ni0.85Se/CP and NiSe–Ni0.85Se/CP.156 The optimized NiSe–Ni0.85Se/CP exhibited a remarkably higher catalytic activity for both OER and HER compared to single-phase catalysts. DFT calculations confirm that H and OH− can be more easily adsorbed on NiSe–Ni0.85Se than on NiSe and Ni0.85Se catalysts. Tian et al. have prepared well-dispersed Ni0.85Se nanoparticles on nitrogen-doped graphene oxide (NGO), resulting in high HER performance.66 Among all the substrates, Ni foam (NF) is a common choice to fabricate nickel selenides for electrochemical water splitting. Zhu et al.157 have synthesized NiSe2, NiSe, and Ni3Se2 on NF via a facile electrodeposition route at a deposition potential of −0.35, −0.46, and −0.60 V, respectively. Such three phases of nickel selenide nanostructures exhibited clear phase-dependent electrocatalytic activities for HER and OER. Typically, NiSe2/NF required overpotentials of only 104 mV and 279 mV to obtain the current densities of 10 mA cm−2 for HER and 20 mA cm−2 for OER, respectively. For the HER, the activity order is NiSe2 > NiSe > Ni3Se2 and for the OER, the order is NiSe2 > Ni3Se2 > NiSe. The phase engineering and utilization of NF verified the importance of the synergistic effects of electrical conductivity and ECSA of the nickel selenide based electrodes. Shi et al. have developed a solvothermal route to synthesize Ni3Se2 rich-grain-boundary nanowire arrays on nickel foam (Ni3Se2/NF). The optimized Ni3Se2/NF-0.4 exhibited superior OER and HER activity. Furthermore, it has been demonstrated as an efficient water electrolyzer, achieving a water-splitting current density of 10 mA cm−2 at a cell voltage of 1.62 V.158
![]() | ||
Fig. 13 SEM images of (a) and (b) NiO/CP and (c) and (d) the NiSe–Ni0.85Se/CP hybrid. (e) Polarization curves of NiSe/CP//NiSe/CP, Ni0.85Se/CP//Ni0.85Se/CP, NiSeNi0.85Se/CP//NiSe-Ni0.85Se/CP, and RuO2/CP(+)//Pt–C/CP(−) pairs in a two-electrode configuration. (f) Multistep chronopotentiometric (CP) curve of the NiSe–Ni0.85Se/CP//NiSe–Ni0.85Se/CP electrolyzer at varying current densities. (g) Chronopotentiometric curves of NiSeNi0.85Se/CP//NiSe–Ni0.85Se/CP. (h) Schematic of the overall water splitting setup system using NiSe–Ni0.85Se/CP as both cathode and anode.156 Reproduced with permission from ref. 156. Copyright 2018, Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim. (i) Formation mechanism of the macroporous MoSe2–rGO composite with empty nanovoids created by applying a pilot-scale spray-drying process. Morphologies of the MoSe2–rGO–L composite powders: (j) and (k) SEM images and (l) elemental mapping images.160 Reproduced with permission from ref. 160. Copyright 2017, Elsevier B.V. All rights reserved. |
For MoSe2, MoSe2/graphene hybrids have gained significant attention due to their potential in improving electron transfer efficiency. Yang et al. successfully prepared MoSe2/RGO hybrids, which exhibited a lower overpotential of 0.05 V vs. RHE compared to previously reported MoS2/graphene hybrids. This improvement can be attributed to the enhanced 3-D electrical contact and the presence of abundant folded edges in the MoSe2/RGO hybrids.159 In addition, the macroporous MoSe2–rGO composite with empty nanovoids was prepared through one-step post-treatment of spray-dried powders as shown in Fig. 13(i)–(l). The MoSe2–rGO–M composite (20 wt% rGO) exhibited a current density of 10 mA cm−2 at a small overpotential of 0.21 V, and a Tafel slope of 57 mV dec−1.160 The outstanding HER performance was due to the synergistic effects of the rGO sheets and MoSe2 nanocrystals, as well as the crumpled structure with empty nanovoids, providing a large number of active sites and facilitating high accessibility for electrolyte ions. Furthermore, the presence of rGO effectively prevented the stacking and growth of few-layered MoSe2. Carbon nanotubes (CNTs) are another commonly used matrix for catalysts. Liu et al. reported a unique hierarchical nanostructure, where few-layered MoSe2 nanosheets are perpendicularly grown on CNTs.161 This hybrid structure exhibited excellent catalytic activity for HER, with a low onset potential of −0.07 V vs. RHE, and a small Tafel slope of 58 mV dec−1. The one-dimensional CNTs effectively prevented the agglomeration and restacking of MoSe2 nanosheets, maximizing the number of exposed active sites and promoting electron transfer efficiency. Two types of substrates are adopted to regulate the catalytic performance; Kang et al. combined MoSe2-reduced graphene (rGO) with CNTs to prepare highly porous MoSe2–rGO–CNT powders by a spray pyrolysis process.162 MoSe2–rGO–CNT exhibited an overpotential of 0.24 V at a current density of 10 mA cm−2 and a Tafel slope of 53 mV dec−1. The high aspect ratio of CNTs provided a porous, spherical backbone that minimized the stacking of rGO nanosheets, while the synergistic effect of CNTs and rGO minimized the growth of MoSe2 nanocrystals. Overall, the synergy of rGO sheets, CNTs, and MoSe2 nanocrystals, along with the unique porous structure of the material, contributed to its exceptional HER properties. In a similar vein, Jia et al. synthesized a MoSe2/reduced graphene oxide/polyimide powder composite, which exhibited improved HER properties compared to unmodified MoSe2.163 Nitrogen doping into the substrate can further enhance the catalytic performance of substrates. Zhang et al. synthesized MoSe2/N-doped reduced graphene oxide (NG) composites.55 The composites consisted of dispersed nanoclusters of MoSe2 nanosheets on folded NG nanosheets. Electrochemical measurements indicated that increasing the N/C ratio initially improved the activity of MoSe2/NG. However, at high N/C ratios, an energy barrier suppressed electron transfer from NG to MoSe2, leading to reduced activity. At low N/C ratios, the impact of the interfacial energy barrier between MoSe2 and NG could be negligible, allowing effective electron transfer. Consequently, N-doped RGO/MoSe2 with an intermediate N/C ratio exhibited the highest activity, with an overpotential of −0.229 V (vs. RHE) at 10 mA cm−2 and a Tafel slope of 78.45 mV dec−1. Other bio-substrates, such as bacterial cellulose-derived carbon nanofibers (CBC), have also been utilized.76 Liu et al. employed CBC nanofibers as a substrate to facilitate the uniform growth of few-layered MoSe2 nanosheets. The carbonized bacterial cellulose provided a 3D network for electrolyte penetration and accelerated electron transfer, resulting in fast hydrogen evolution kinetics with an onset overpotential of 91 mV and a Tafel slope of 55 mV dec−1 in acidic media. Therefore, incorporating biotemplate materials offered a novel approach to develop high-performance catalysts for HER, simultaneously increasing active sites and conductivity.
Dual-native vacancies of Se and Mo can be introduced in 2H-MoSe2 nanosheets, which alter the catalytic sites and improve the electrochemical performance.45 These vacancies can be generated in CVD grown TMDs by controlling precursor vapor, pressure, and growth temperatures during synthesis. First-principles calculations revealed that both Se and Mo vacancies alter the catalytic sites, optimizing the basal plane and edges with the optimal free energy for HER. Additionally, the vacancies increase the number of gap states and electrons near the Fermi level, improving electron transport efficiency. Experimental results demonstrated that MoSe2 nanosheets with a significant amount of dual-native vacancies exhibited a low overpotential and good stability for 20 h I–t testing. These findings indicate the effectiveness of defect engineering in enhancing catalytic activity.
As shown in Fig. 14, the controllable defect modulation can be achieved by engineering selenium vacancies in WSe2 monolayer nanosheets through annealing.164 The introduction of Se-vacancies created more active sites on the basal plane, allowing for favorable hydrogen adsorption. First-principles calculations were conducted to confirm the HER activity of the Se-vacancy sites. Increasing the number of Se vacancies enhanced neutral hydrogen adsorption, reduced the energy barrier, and lowered the overpotential.
![]() | ||
Fig. 14 (a) Schematic illustration of the top (upper panel) and side (lower panel) views of WSe2 with Se-vacancies on the basal planes, where the Se-vacancies serve as the active sites for hydrogen evolution. (b) HER free energy in the Se-vacancy range of 0–25%.164 Reproduced with permission from ref. 164. Copyright 2016, The Royal Society of Chemistry. |
Yin et al. introduced a novel method to simultaneously modulate the phase structure and unsaturated defects in partially crystalline 1T-MoSe2.116 The synthesis process of 1T-MoSe2 involved the reductant NaBH4, which played a key role in forming the 1T phase. By conducting the synthesis at lower reaction temperatures, a disordered structure with numerous active defects can be obtained. The combination of the high conductivity of the 1T phase and defective surface led to a remarkable catalytic activity improvement for HER.
Overall, this part has highlighted and summarized the significance of defect engineering in enhancing the catalytic performance of electrocatalysts. The potential mechanism may be the modification of the catalytic sites and improvement of electron transport efficiency.
![]() | ||
Fig. 15 (a) Schematic representation of the synthesis of the mesoporous CoSSe catalyst. (b) Schematic structural representation of H adsorption on different {001} surface elements: (I) on Co of CoS2, (II) on S of CoS2, (III) on Co of CoSSe, and (IV) on S of CoSSe.167 Reproduced with permission from ref. 167. Copyright 2019, American Chemical Society. (c) Schematic illustration outlining the fabrication of the W(SexS1−x)2 nanoporous architecture. (d) and (e) TEM images of the W(SexS1−x)2–15 NPA. (f) HRTEM image of the region marked with a yellow dashed box in panel (e). The light blue boxes represent lattice mismatch and defects. Fast Fourier transform analysis performed on the regions marked by (g) orange and (h) light green dashed boxes.168 Reproduced with permission from ref. 168. Copyright 2017, American Chemical Society. |
For WSe2, heteroatom doping could engineer the phase to metallic 1T and fabricate defects simultaneously. For example, as shown in Fig. 15(c)–(h), Liang et al. developed the W(SexS1−x)2 nanoporous architecture (NPA) by substituting Se with S.168 This ternary W(SexS1−x)2 NPA exhibited significant advantages for high-efficiency HER. The nanoporous morphology provided more electrochemically active sites for water splitting. The lattice mismatch and disordering in the mix-phased W(SexS1−x)2 film introduced additional defects, enhancing the catalytic activity. The formation of a conducting 1T phase in the strained W(SexS1−x)2 facilitated electron transfer during catalysis. As a result, the as-prepared W(SexS1−x)2 NPA achieved an onset overpotential of 45 mV and a Tafel slope of 59 mV dec−1.
For NiSe2, the introduction of nitrogen (N) in conductive NiSe2 can promote the redistribution of charge densities and enrich the active sites, effectively enhancing its HER activity. Wang and co-workers developed a self-supported 3D structured nitrogen-doped nickel selenide (N-NiSe2) by in situ growing NiSe2 on nickel foam under an ammonia atmosphere (N-NiSe2/NF).169 This N-NiSe2/NF electrode exhibited excellent electrocatalytic activity for the HER across a wide pH range (pH = 0.3, 7, and 14). In 1.0 M KOH, the N-NiSe2/NF electrode demonstrated a small overpotential of 86 mV to achieve a current density of 10 mA cm−2. The electrocatalytic effects of Mo in pure NiSe for the HER and OER have been explored.170 Coral-like structured Ni1−xMoxSe has been prepared. The distinct morphology of the Ni0.5Mo0.5Se material exhibited enhanced electrochemical activity, following the Volmer–Heyrovsky mechanism. This work investigated the influence of the electrocatalytic properties of Mo in pure NiSe towards the HER and OER. Sun et al. synthesized an atomically thin Ni–Se–S based hybrid nanosheet (NiSe1.2S0.8).171 Theoretical results revealed that the hybrid electrocatalyst with sulfur (S) incorporation exhibited optimal adsorption free energy of hydrogen (ΔGH*).
For MoSe2, various strategies and doping techniques have been developed to enhance the catalytic activity of MoSe2 for the HER. These strategies include doping other elements (N, B, S) to narrow the bandgap of MoSe2 and increase active sites on the basal planes.10,47,172,173 For instance, as shown in Fig. 10, Tu et al. synthesized high-quality N-doped MoSe2/TiC-C shell/core arrays via a facile hydrothermal plus annealing process, combining the phase and morphology engineering strategies together.174 N-doping not only decreased the band gap of MoSe2, but also reduced the energy barrier promoting easier phase transition from 2H to 1T. A systematic computational study revealed that B dopants can well function in lowing the formation energy in MoSe2, suggesting the feasibility of B doping in activating the catalytic behavior of MoSe2. Experiments have demonstrated that the doping of B into MoSe2 could lead to the narrowed band and increased electrical conductivity, largely facilitating electron movement and charge transfer. Especially with the doping of more B atoms into the MoSe2 lattice, the bands of MoSe2 moved closer to the Fermi level, activating the conductivity of the basal plane.47 The incorporation of S into MoSe2 has also been reported, such as high-percentage 1T-phase, single-layer ternary nanodots (MoSSe) with high-density active edge sites.175 The synthesis of MoSSe nanodots with high-density active edge sites promoted fast electron/charge transfer and resulted in an alloying effect between S and Se atoms, as well as the presence of Se vacancies on MoSSe nanodots. These factors contributed to the enhancement of catalytic activity on the basal plane. The MoSSe ternary nanodots exhibited significantly improved electrochemical HER performance, characterized by a low overpotential of −140 mV at a current density of 10 mA cm−2, a Tafel slope of 40 mV dec−1, and excellent long-term durability. Additionally, doping with Zn69 and introducing Pt nanoparticles176 have been investigated as effective strategies to improve the catalytic activity of MoSe2. DFT calculations showed that the ΔGH* of the Zn-doped site was −0.02 eV, and the adjacent Se site changed from 1.93 eV in pure MoSe2 to −0.15 eV. The final Zn-doped MoSe2 possessed the smallest onset potential (182 mV) and the lowest Tafel slope (58 mV dec−1). The final MoSe2–Pt catalyst exhibited an onset overpotential of 100 mV and a Tafel slope of ∼85 mV dec−1.
Nickel–cobalt selenides (Ni–Co–Se) or their compounds with other non-metal materials are commonly used as multi-metal catalysts. Cobalt selenides and nickel selenides both can be regarded as high performance catalysts for HER and OER. The combination of active Co and Ni can further improve electrochemical activity and kinetics by modifying the electronic structure. The combined contribution of Ni and Co elements to the electronic structure of bimetallic selenides reveals the strong interaction of bimetals toward enhanced electrical conductivity. The synergistic effect between cobalt and nickel selenides could further improve the conductivity and activity, which is believed to contributed to their electrocatalytic activity for the OER. This is because OER is described as a kinetically sluggish process involving a multistep proton-coupled electron transfer. The intrinsically metallic Co0.85Se can be further promoted by Ni doping. The metallic nature of Co0.85Se, which is in sharp contrast to the semiconducting nature of other cobalt selenides such as CoSe and Co0.85Se, makes it useful as an electroactive material or support material to non-conductive OER compounds.
Metallic Co0.85Se and (Ni,Co)0.85Se nanotube arrays were synthesized on a carbon fabric collector (CFC) for the OER.177 The (Ni,Co)0.85Se nanoarrays exhibited superior performance in an alkaline medium. This improvement can be attributed to the better electrical conductivity and higher defect concentrations, which serve as active sites for catalysis. Nanostructured Ni-doped cobalt selenides, with their excellent conductivity, corrosion resistance, and promising OER activity, can be utilized as advanced OER catalysts or as backbone materials for integrating insulating OER active materials in hybrid catalysts. A Co-doped nickel selenide (a mixture of NiSe2 and Ni3Se4)/C hybrid nanostructure supported on Ni foam (Co–Ni–Se/C/NF) was prepared using the organic framework of zeolitic imidazolate framework-67 (ZIF-67) as a precursor.178 The as-obtained catalyst exhibits excellent catalytic activity for the OER at a potential of 275 mV at 30 mA cm−2, as well as efficient catalysis for the HER at a potential of 90 mV at 10 mA cm−2. Furthermore, this hybrid nanostructure demonstrated good durability and achieved current densities of 10 and 30 mA cm−2 at potentials of 1.6 and 1.71 V, respectively, when used as both cathode and anode for overall water splitting in basic media.
Other catalysts based on Co-doped nickel selenides have been developed for the HER and OER in water splitting. Each study has presented a different synthesis method and highlighted the performance and characteristics of the catalysts. Ao et al. synthesized yolk-shelled Ni–Co–Se nanocages on carbon fiber paper (Y–S Ni–Co–Se/CFP), which exhibited remarkable electrocatalytic activity and long-term stability for both HER and OER. This work provides another way to design MOF-derived functional materials with interesting nanostructures for diversified applications.179 Zhang et al. regulated the electronic property and surface structures of cobalt-doped Ni0.85Se chalcogenides and found that Co0.1Ni0.75Se showed the highest catalytic activity for HER.180 They further improved its performance by using r-GO as a support (Co0.1Ni0.75Se/rGO), which exhibited an overpotential of 103 mV at 10 mA cm−2 and a Tafel slope of 43 mV dec−1 and the current tended to be stable for 30 h i–t testing. Gao et al. designed CoNiSe2 heteronanorods decorated with layered-double-hydroxide (LDH) nanosheets on nickel foam (CoNiSe2@CoNi-LDHs/NF), presenting enhanced HER kinetics due to balanced water dissociation and H desorption at the selenide–LDH interfaces.72 DFT calculations show that after Co-doping, CoNiSe2@Co-LDHs and CoNiSe2@Ni-LDHs presented a visibly downshifting d-band center of −2.91 and −2.84 eV, respectively, compared to bare CoNiSe2 (−2.74 eV), indicating the electron transfer through CoNiSe2–LDH interfaces. In addition, the H2O chemisorption energy on the interface was also reduced to −1.83 and −1.55 eV after being decorated by Co- and Ni-LDHs. The CoNiSe2@CoNi-LDHs/NF nanocomposites showed the best activity, with an overpotential of 106 mV at 10 mA cm−2, a Tafel slope of 74 mV dec−1 and a high value of ECSA of 186.2 cm2. Wang et al. used Prussian blue analogues (PBAs) as precursors to synthesize penroseite (Ni,Co)Se2 nanocages anchored on 3D graphene aerogel ((Ni,Co)Se2–GA). Consequently, (Ni,Co)Se2–GA showed an overpotential of 128 mV at −10 mA cm−2 and a Tafel slope of 79 mV dec−1 for water splitting with a low voltage of 1.6 V at 10 mA cm−2 and possessed outstanding durability.181 Yu et al.182 designed novel NiCoSe2 nanosheet networks grown on carbon cloth (CC) with crimped nanosheet configuration via an electrodeposition technique. NiCoSe2 exhibits superior electrocatalytic performance and robust durability toward overall water splitting, especially toward oxygen evolution reaction (OER) with a low overpotential of 255.8 mV to deliver 10 mA cm−2 current density. Xiao et al. have prepared (Ni, Co)0.85Se NSAs/NF for super overall water splitting.183 The NF substrate transforms from a hydrophobic surface to a hydrophilic one after modification with porous structure of (Ni,Co)0.85Se NSAs, which perpendicularly grown on the NF. The catalysts grown on the substrate in situ are more likely to adsorb droplets and promote traps of electrolyte ions and access to active sites. The as-obtained (Ni,Co)0.85Se NSAs exhibit a low overpotential of 169 mV at a current density of 10 mA cm−2 for the HER and an overpotential of 287 mV at a current density of 20 mA cm−2 for the OER, and superior performance toward overall water splitting with a cell voltage as low as 1.65 V at a current density of 10 mA cm−2. A Ti plate can be used as the conductive substrate; Liu et al.184 have reported Co0.13Ni0.87Se2/Ti on a conductive Ti plate for both HER (10 mA cm−2, 64 mV) and OER (100 mA cm−2, 320 mV) in strongly basic media (1.0 M KOH). A voltage of only 1.62 V is required to drive 10 mA cm−2 for the two-electrode alkaline water electrolyzer. Chen et al.185 have prepared multilayered CoNiSe2 nanorods (NRs) that grow radially from the substrate to form hierarchical sea urchin-like microstructures. The overpotentials required to deliver a current density of 100 mA cm−2 are as low as 307 and 170 mV for the OER and HER, respectively; for water splitting reaction, the overpotential required is 1.591 V to achieve a current density of 10 mA cm−2. These studies demonstrated the synthesis of various catalysts with improved performance for HER and OER, highlighting the importance of composition, structure, and support materials in designing efficient catalysts for applications.
Nickel-based selenide materials are promising candidates for oxygen evolution reaction due to their low cost and excellent performance. The incorporation of Fe into the nickel-based selenide can facilitate charge transportation, expand electrochemically active surface and enhance desorption of oxygen. In recent years, NiFe hybrid selenides have been proposed as exciting OER catalysts owing to the sufficient collaboration of nickel and iron. Based on this, various types of NiFe hybrid selenides have been developed as efficient catalysts for the OER in water splitting. One study has reported a nanostructured nickel iron diselenide (NixFe1−xSe2), which acted as a precursor and then in situ converted into NixFe1−xSe2-derived oxide (NixFe1−xSe2-DO) under a stable current density of 5 mA cm−2. As shown in Fig. 16(a)–(e), NixFe1−xSe2-DO gave an ultra-low overpotential (195 mV at 10 mA cm−2). This work emphasized the importance of identifying the active species of OER catalysts.186 The polyvinyl-pyrrolidone (PVP)-decorated Ni–Fe diselenide hollow nanoparticles (P-NFSHPs) with strong hydrophilicity have been explored as an electrocatalyst for OER. The strong polarity of lactam groups in PVP introduced a highly hydrophilic surface to the selenide catalyst. The as-prepared P-NFSHSs with enhanced wettability displayed enhanced OER performance with a low overpotential of 255 mV (vs. RHE) at 10 mA cm−2, a low Tafel slope of 56 mV dec−1 and decent long-term stability (only 13 mV degradation after 12 h electrolysis).187 3D porous Ni-FeSe with a rose-like microsphere architecture directly grown on Ni foam has also been developed with a two-step pathway.188 The unique 3D mesoporous rose-like morphology facilitated mass and electron transport as well as O2 bubble release. The resulted Ni0.76Fe0.24Se45 exhibited superior OER performance. Du et al. have fabricated hierarchical (Ni,Fe)3Se4 ultrathin nanosheets which possess a low overpotential of 225 mV and a small Tafel slope of about 41 mV dec−1 for OER, relying on abundant and accessible catalytically active sites,189 facile charge transfer and a high specific surface area.
![]() | ||
Fig. 16 (a) Illustration of the synthetic route to Co–Ni–Se/C/NF. (b) Polarization curve of Co–Ni–Se/C/NF in a three-electrode system in a wide potential window (−0.28 to 1.65 V vs RHE). (c) Polarization curves of RuO2‖Pt/C, Co–Ni–Se/C/NF and NF. (d) Chronopotentiometric curve for Co–Ni–Se/C/NF under a constant current density of 50 mA cm−2 for 100 h. (e) Energy efficiency of the Co–Ni–Se/C/NF electrolyzer as a function of current density.186 (f) Schematic illustration of the synthesis strategy of the heterostructures. SEM images of (g) NiCH@CC, (h) NiFe-PBA@NiCH@CC, and (i) NiFeSe@NiSe|O@CC. Typical low-magnification SEM images of (j) NiFe-PBA@NiCH@CC. Electrochemical HER and OER performances of NiFeSe@NiSe|O@CC along with the NiFeSe nanocubes, NiSe|O@CC, CC, Pt/C@CC, and RuO2@CC for comparison. (k) HER polarization curves in 1 M KOH. (l) Corresponding Tafel plots of the electrocatalysts in (k). (m) Time dependence of current density under static overpotential showing the durability of the electrocatalysts over 50 h. (n) OER polarization curves in 1 M KOH. (o) Corresponding Tafel plots of the electrocatalysts in (n). (p) Time dependence of current density under static overpotential showing the durability of the electrocatalysts over 50 h.191 Reproduced with permission from ref. 191. Copyright 2018, Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim. |
Fe-doped NiSe (a mixture of hexagonal NiSe and orthorhombic NiSe) nanorod/nanosheet hierarchical arrays have been formed, which exhibited remarkable OER activity and decent HER activity.190 The self-templated pseudomorphic selenization method has been adopted for transforming nickel iron Prussian blue analogue (NiFe-PBA) and nickel carbonate hydroxide (NiCH) precursors into a porous and defect-enriched nickel iron selenide hybrid on carbon cloth (NiFeSe@NiSe|O@CC). The final product exhibited highly efficient and stable electrocatalytic activity toward HER and OER, as well as overall water splitting in alkaline media with a low overpotential of 1.56 V, as shown in Fig. 16(f)–(p).191 Wang et al.192 reported the porous nickel–iron bimetallic selenide nanosheets ((Ni0.75Fe0.25)Se2) on carbon fiber cloth (CFC) by selenization of the ultrathin NiFe-based nanosheet precursor. The as-prepared three-dimensional oxygen evolution electrode exhibits a small overpotential of 255 mV at 35 mA cm−2 and a low Tafel slope of 47.2 mV dec−1 and maintains high stability during a 28 h measurement in alkaline solution. Umapathi et al.193 have reported FeNi2Se4 nanoparticles supported on nitrogen doped reduced graphene oxide (FeNi2Se4–NrGO) as an efficient catalysis for oxygen evolution under alkaline conditions. While FeNi2Se4 nanoparticles themselves show good catalytic activity for water oxidation, the constructed hybrid nanocomposite with NrGO as the supporting matrix shows enhanced catalytic activity with a small overpotential of 170 mV @ 10 mA cm−2, a small Tafel slope of 62.1 mV per decade, and high current density. This study verified the effect of anion coordination on the catalytic performance, as well as the synergistic role of nanoscale interactions between the catalyst particles and graphene matrix in enhancing catalytic activity.
Other bimetallic selenides have been reported for water electrolysis; for example, a Fe–Co-based PBA precursor was designed by self-assembly and subsequently selenized to (Fe–Co)Se2. The product of (Fe–Co)Se2 only required an overpotential of 251 mV to drive a current density of 10 mA cm−2 in the OER system, and 90 mV was required at a current density of 10 mA cm−2 toward the HER activity, along with a low Tafel slope of 47.6 and 58.7 mV. Furthermore, the catalytic performance of Se-enriched Co1−xFexSe2 catalysts on nickel foam can be adjusted according to the Co/Fe ratios in the catalyst. The Co0.4Fe0.6Se2 nanosheets not only exhibited superior OER performance, at a current density of 10 mA cm−2 with an overpotential of 217 mV, but also possessed ultrahigh durability with a dinky degeneration of 4.4% even after a 72 h fierce water oxidation test in alkaline solution.194
Fe, Co, and Ni are common heteroatoms to enhance the electrocatalytic active sites, improve the electronic conductivity, and modulate the intermediates adsorption energy of TMSes. In addition to Fe, Co, and Ni, other transition metal atoms have also been explored to optimize the electrocatalytic properties of TMSes. Zhao et al.195 have prepared Mn-regulated cobalt selenide nanosheets. With the modulation of Mn, the tailored atomic disorder, tuned electronic structure, together with optimized electrical conductivity could be simultaneously achieved in CoMn selenide. The engineered structural and electrical properties led to the effective generation of active species and promoted the reaction rate during the OER process, in accordance with the high catalytic activity. The overpotential of (CoMn)Se2 catalysts was 0.27 V at a current density of 10 mA cm−2, much lower than that of CoSe2 catalysts and the state-of-the-art IrO2. MoCoSex encapsulated in N-doped hollow carbon nanospheres (MoCoSex@NC) has been reported and it exhibited an excellent low overpotential of 60 mV at 10 mA cm−2 and great stability for HER.196 V0.86Co0.14Se2 was synthesized by doping cobalt into vanadium diselenide (VSe2) nanosheets on a carbon cloth (CC), presenting a low overpotential of 230 mV at 10 mA cm−2, and a small Tafel slope of 63.4 mV dec−1.197 Ternary refractory metal selenides (MWSex; M = Fe, Co, Ni and Mn) have been formed by introducing transition metals into tungsten selenide. The catalytic activity differences of ternary metal selenides have been compared, and they were ranked in the order of Fe ≈ Mn, Co < Ni. The addition of Ni and Co into the tungsten selenide made them more active catalysts compared to pristine WSex.198 All these catalysts have the potential for practical catalytic applications in water electrolysis.
Multi-metal selenides generally exhibit superior catalytic performance compared to their monometallic counterparts. This can be attributed to their enhanced electrical conductivity, higher structural stability, and favorable Gibbs free energy of hydrogen absorption (ΔGH*). The ternary selenides of Fe–Cu exhibit reduced OER activity compared to their pure parent compounds. However, with the addition of the Co dopant, all Fe–Co–Cu quaternary selenides exhibited enhanced catalytic activity, with lower overpotential and lower Tafel slopes. The relative amounts of Fe and Cu in the catalysts play a significant role in catalytic activity, with increased amounts of either Fe or Cu leading to improved activity. The optimal catalyst, (Fe0.48Co0.38Cu0.14)Se, required a low overpotential of 256 mV to achieve 10 mA cm−2 and showed a favorable Tafel slope of 40.8 mV dec−1. The enhanced catalytic activity is attributed to electron cloud delocalization and the formation of d-bands among the transition metal sites, as shown in Fig. 17(a)–(d).200 Multi-metal iron–nickel–cobalt selenide (Fe0.37Ni0.17Co0.36Se) has been proved to be superior to the ternary nickel cobalt selenide in terms of conductivity and electrocatalytic activity, with an overpotential of −179 mV at 10 mA cm−2 and a Tafel slope of 115.3 mV dec−1.65 The OER catalytic activity of Co–Ni–Cu selenides in alkaline medium is affected by transition metal doping, which is sensitive to the concentration of Cu in the catalysts. An increase in Cu concentration leads to an increase in activity. However, beyond a certain limit, a higher Cu concentration results in a decrease in catalytic efficiency. The Cu3Se2 structure shows excellent OER catalytic activity, with a low overpotential of 326 mV at 10 mA cm−2. The (Co0.21Ni0.25Cu0.54)3Se2 thin film also exhibited excellent OER catalytic activity (272 mV, 10 mA cm−2). Compared to ternary and binary selenides, quaternary compositions exhibit higher catalytic activity, indicating a synergistic effect of transition metal doping in enhancing catalytic activity, as shown in Fig. 17(e)–(h).201 Furthermore, the catalytic activity of mixed metal selenides comprising various compositions of Fe, Co, and Cu selenides has been systematically investigated.
![]() | ||
Fig. 17 (a) Schematic of combinatorial electrodeposition. (a) Ternary phase diagram for exploring the compositions of the mixed-metal selenide films examined in this work. Crossing vertices represent compositions of the precursor electrolyte with respect to the relative ratio of the corresponding metals. The black circle shows a typical example. (b) Typical experimental set-up for electrodeposition. Contour plots of overpotential η (in units of V) (c) at the onset of OER activity and (d) with a current density of 10 mA cm−2 for the entire Ni–Fe–Co trigonal phase space.200 Reproduced with permission from ref. 200. Copyright 2019, American Chemical Society. Contour plots of onset overpotential (in units of V) (e) and overpotential (in units of V) (f) at a current density of 10 mA cm−2 for the entire Co–Ni–Cu trigonal phase space. The color gradient corresponds to the overpotential measured in volts. (g) Polarization curves of (Co0.21Ni0.25Cu0.54)3Se2 in comparison to the binary selenide films. (h) SEM image of the (Co0.21Ni0.25Cu0.54)3Se2 film and the inset is a higher magnification image.201 Reproduced with permission from ref. 201. Copyright 2019, The Royal Society of Chemistry. |
Self-supported porous crystalline Zn–Ni–Co/Cu selenide microball arrays have been constructed through a series of steps.202 Tubular Cu(OH)2 arrays were first fabricated on carbon cloth and then covered with ultrathin Zn–Co–Ni layered hydroxide to form a core–shell structure. The Zn–Co–Ni layered ternary hydroxides were then deposited around the tubular arrays. Through selenization, the morphology changed from amorphous tubes to highly crystalline microballs, resulting in nanoporous ZnNi0.5Co0.5Se2/Cu1.8Se microballs. These advanced electrode materials possessed high conductivity, large specific surface areas, as well as multiple redox and electrocatalytic active sites for the OER. They exhibited a remarkable catalytic current of 325 mA cm−2 at 1.6 V vs. RHE for OER in alkaline solutions. The catalytic activity of these materials is attributed to the synergistic effect of Ni, Co, Zn, and Cu redox sites within the well-crystallized cubic structure of ZnNi0.5Co0.5Se2. Furthermore, the materials exhibit chemical stability and good mechanical strength, enabling them to withstand undesirable etching and accommodate volume changes during cycling.
Hierarchical trimetallic hydroxides on Ni foam can be selenized to obtain Ni–Co–Fe–Se@NiCo-LDH, exhibiting excellent electrocatalytic performance for overall water splitting in alkaline medium; it required an overpotential of 286 mV to deliver a current density of 100 mA cm−2 for OER and 113 mV at a current density of 10 mA cm−2.203 The catalyst required low overpotentials for both OER and HER. Additionally, the material showed excellent overall water splitting performance with a low cell voltage of 1.55 V at 10 mA cm−2. DFT calculations indicated that the presence of Co-NiSe2 accelerated hydrogen production kinetics, while Fe7Se enhanced material conductivity. The synergistic effect of these components in the Ni–Co–Fe–Se@NiCo-LDH catalyst led to enhanced hydrogen production activity.
The mixed metal selenides generally show improved catalytic activity compared to their pure parent compounds. The addition of transition metal dopants, such as Cu, Co, Fe, Ni, and Zn, enhanced the catalytic activity by promoting electron cloud delocalization, improving charge transport resistance, and increasing film conductivity. Proper transition metal doping can further increase the quantity of active sites and modulate the reaction kinetics by regulating OH adsorption on the electrocatalyst surface and modulating the local electron density near the active sites. Furthermore, the highly occupied d-levels improve the electrical conductivity within the matrix to accelerate charge transfer on the electrocatalyst surface. The synergistic effects arising from the combination of different metal selenides also contribute to the overall improved catalytic activity. These findings offer valuable insights into the development of catalytic properties, which hold significant practical implications for application in electrocatalytic water splitting.
The scalable and cost-effective production of clean and sustainable fuel alternatives, specifically hydrogen through electrochemical water decomposition, is an urgent goal due to the increasing energy crisis and environmental pressure. TMSes have emerged as promising electrocatalysts for water splitting, as they are abundant and can replace noble metal compounds. However, the performance of TMSe-based electrocatalysts, especially in terms of high-current-density and long-term stability, requires further improvement to meet the demands of industrial applications. Therefore, there are many potential directions for future research in this field.
TMSes have shown potential as electrodes in energy utilization and water splitting devices, but their applications should be continuously extended. Bandgap engineering of TMSes can be applied for photocatalysis, photodetectors, and photo-electrochemical sensors, which can be further regulated by heteroatom doping, atomic layer regulation, and heterojunction constructing. Appropriate modulation of the electronic structure of TMSes may render them desirable for applications such as CO2 reduction, nitrogen reduction, environmental control processes, sensors, devices (ion batteries, supercapacitors), and dye-sensitized solar cells.
This journal is © the Partner Organisations 2024 |