2D transition metal-based phospho-chalcogenides and their applications in photocatalytic and electrocatalytic hydrogen evolution reactions

K. Pramoda *a and C. N. R. Rao *b
aCentre for Nano and Material Sciences, Jain (Deemed-to-be University), Jain Global Campus, Kanakapura, Bangalore, Karnataka 562112, India
bNew Chemistry Unit, School of Advanced Materials and International Centre for Material Science, Jawaharlal Nehru Centre for Advanced Scientific Research (JNCASR), Jakkur P. O., Bengaluru, 560064, India. E-mail: cnrrao@jncasr.ac.in

Received 17th March 2023 , Accepted 30th June 2023

First published on 10th July 2023


Abstract

The utilization of hydrogen (H2) as a renewable substitute for fossil fuel can mitigate issues related to energy shortage and associated global warming. The generation of H2via water splitting by photo and electrocatalytic means is of significant importance. The employment of a semiconductor catalyst reduces the high energy barrier (237 kJ mol−1) associated with a water-splitting reaction to yield H2 and O2. Sustainable H2 production demands the utilization of noble-metal-free, efficient, and stable catalysts over a wide pH range. In recent times, layered transition metal thio(seleno) phosphates (MPX3, X = S, Se) are reported to be highly efficient for H2 evolution reaction owing to their earth-abundance, rich active sites, wide spanned band gap of 1.2 to 3.5 eV, and high chemical stability. In this perspective, advances in 2D monometallic and bimetallic MPX3 compounds are reviewed comprehensively from the viewpoint of water splitting, especially the hydrogen evolution reaction (HER). This study includes the composition, structural engineering, heterostructure, and hierarchically structural design in enhancing the HER activity of MPX3. Computational results providing insights into the intrinsic photo and electrocatalytic HER activity of 2D MPX3 are presented. Finally, the challenges and opportunities in further improving MPX3 activity towards HER and associated catalysis reactions are discussed.


1. Introduction

The massive utilization of non-renewable fossil fuels to meet ever-increasing global energy requirements is leading to fuel scarcity and CO2 emission at an alarming rate gives rise to global warming.1–8 Therefore, establishing global-scale, renewable, efficient, and carbon-free emission combined fuel-combustion technology is of significant importance.1–4 Fuel cell technology is an environmentally-safe route of generating energy than the other existing methods.5–7 The fuel cell utilizing hydrogen (H2) and oxygen (O2) as fuels at the anode and cathode, respectively, will be the suitable one amongst the numerous types of fuel cells because it gives only water (H2O) as the product.8–10 The main goal is to generate H2 from renewable sources, ideally H2O. Water can be decomposed into H2 and O2 through photochemical, electrochemical, and photoelectrochemical strategies by utilizing solar, electrical energy, and a combination of solar-electricity energy, respectively.11–14,19 The main challenge is to generate H2 from renewable sources, preferably H2O. But, water splitting image file: d3ta01629c-t1.tif is thermodynamically an uphill reaction that involves a large Gibbs free energy change of 237 kJ mol−1.15–18 With the help of an appropriate catalyst one can decrease the energy barrier for water splitting and increase the H2 evolution rate. Platinum or platinum group materials (PGMs) are regarded as the best catalysts for electrochemical H2 evolution reaction (HER).20–22 However, high-cost, stability issues under various electrochemical conditions and intolerance to catalytic poisoning severely limit their applicability as HER catalysts. In the last decade, there is a major thrust into exploiting low-cost, earth-abundant, and stable catalysts, particularly the non-precious metal catalysts.23–26 Therefore, non-noble 2D metal catalysts, such as transition metal chalcogenides (TMDCs),27–30 carbides,31 nitrides,32 and phosphides,33 have been investigated for the potential replacement of the noble metal catalysts for water splitting owing to their fascinating optical and electrical properties. It has been reported that the bulk MoS2 is inactive for the hydrogen evolution reaction (HER), whereas, mono or few-layer MoS2 shows excellent HER performance.34,35 The incorporation of Co and Ni into the MoS2 matrix was reported to enhance the electrocatalytic activity significantly.36–38 While alloy phases, MoSxSe1−x, wherein Se in combination with S, permit tuning of HER activity.39,40

Akin to TMDCs, 2D transition metal phosphides (TMPs) are reported to show bifunctional properties as HER and oxygen evolution reaction (OER) catalysts and also display long-term stability in a wide pH range.41–44 Theoretically, it is predicted that the phosphorous atom of TMPs acts as an active site for intermediate hydrogen adsorption and desorption during water splitting. Jia et al.45 compared the Gibbs free energy change (ΔGH*) of hydrogen adsorbed, which is a main intermediate state in HER, on to the different active sites MoP, MoS2, and Mo2C catalysts by density functional theory (DFT) calculations (Fig. 1a and b). According to the Sabatier principle, too strong or too weak adsorption is not favorable for the water reduction reaction. The ΔGH* values for Mo-terminated MoP, MoS2, and Mo2C were calculated as 0.16, 0.19, and 0.24 eV, respectively. These results indicate that the P atom in MoP plays the equivalent role as S of MoS2 and C in Mo2C. Further, a subtle smaller ΔGH* value for MoP as compared to MoS2 and Mo2C, indicates that MoP possesses higher catalytic activity. In another study, Miguel et al.46 using DFT calculations demonstrated that the ternary pyrite-type cobalt phosphosulfide can be an efficient catalyst for both photocatalytic and electrocatalytic HER since the cobalt octahedra contain more-electron donating P2− sites compared to S2−, owing to thermoneutral H+ adsorption at the active sites. The cubic crystal structure of CoPS can be envisaged as Co3+ octahedra surrounded by dumbbells with a uniform arrangement of P2− and S2− ligands (Fig. 1c and d). Further, DFT calculations revealed that ΔGH* on the (100) surface of CoPS at the Co site is more favorable compared to the CoS2 constituent. As shown in Fig. 1f–i, after spontaneous adsorption of H+ at open P sites, adjacent Co3+ atoms are reduced to Co2+, which brings ΔGH* value for cobalt near the thermoneutral value. Subsequent H+ adsorption at Co2+ sites induces the oxidation of Co2+ to Co3+ as shown in Fig. 1f–i. Such synergistic interaction is not realized in the case of CoS2 since Co3+ sites are energetically more favorable for H+ adsorption than open S2− ligands. In addition, Miguel et al.46 showed that CoPS nanostructures display electrocatalytic HER activity comparable to the Pt/C catalyst with a current density of 10 mA cm−2 at an overpotential as low as 48 mV and also shows exceptional stability. The Pt/C-like HER activity of CoPS inspired other researchers to study MPX3 systems containing Mn, Fe, Ni, Cd, etc. metals. But, some of the 2D family MPX3 systems such as MnPS3, FePS3, and NiPS3 showed marginal electrocatalytic H2 evolution activity due to low intrinsic electronic conductivity.47 Coupling of MPX3 with conducting graphene47 and doping of Co into MnPS3 and NiPS3 structures was reported to improve HER activity.48,49 Doping modifies the valence band and conduction edges and improves the electron transport rate. Further, MoS2 in combination with MPX3 facilitates hydrogen spillover during HER from MoS2 edge sites to TMPCs.50,51


image file: d3ta01629c-f1.tif
Fig. 1 (a) Schematic depiction of the layered structure of 2D molybdenum phosphide (MoP), molybdenum disulfide (MoS2), and molybdenum carbide (Mo2C). (b) Gibbs free energy diagram of the H2 evolution for MoP, MoS2, and Mo2C at different sites. Reprinted with permission.45 Copyright 2017, American Chemical Society. (c and d) Crystal structures of CoS2 and CoPS systems. (e) Free energy diagrams of the H2 evolution for CoS2 and CoPS at different sites. Structural demonstration of hydrogen adsorption (H*) at the Co site on (100) plane of CoS2 (f), at the Co site (g), P site (h), and at the P site after H* at the Co site (i) on (100) plane of CoPS. Reprinted with permission. Copyright 2015, Springer nature.46

As discussed in prior sections, water splitting is thermodynamically energy-intensive as it demands a relatively large free energy of 237 kJ mol−1 to produce one mole of H2 and O2 each.15,19 To overcome this large energy barrier, researchers are utilizing semiconductor photocatalysts for water splitting, which can mimic the natural photosynthesis process involving the following two half-reactions.52,53

Water reduction: 2H+[thin space (1/6-em)]+[thin space (1/6-em)]2e → H2, ΔEo = −0.41 V

Water oxidation: 2H2O → O2[thin space (1/6-em)]+[thin space (1/6-em)]4H+ + 4e, ΔEo = +0.82 V

image file: d3ta01629c-t2.tif

The high Gibbs free energy barrier for water redox reactions can be reduced by employing semiconductors as photocatalysts and, therefore, the H2 evolution rate can be improved. The standard reduction potentials for H+/H2 reduction and oxidation potential for H2O/O2 are estimated to be −0.41 and +0.82 V (vs. NHE), respectively. To achieve overall water splitting by utilizing a semiconductor, the conduction band minimum (CBM) energy of the semiconductor should be higher than the H+/H2 reduction potential and its valence band maximum (VBM) energy must be lower than the H2O/O2 oxidation potential.54–56 The water splitting utilizing semiconductor photocatalyst comprises three essential steps: (1) absorption of solar energy by the semiconductor; (2) generation and mobilization of electron–hole pair from bulk to the active sites of the semiconductor; (3) water reduction (H+/H2) and oxidation (H2O/O2) reactions by photogenerated electron and hole pair at the conduction band (CB) and valence band (VB) of the semiconductor, respectively (Fig. 2). As shown in Fig. 2a and b, water splitting using semiconductors could be a one-step or two-step approach based on whether a single component is utilized or a combination of two or more semiconductor systems. The one-step route involves the photoexcitation of the semiconductor and water reduction and oxidation reaction by the electron and holes at the CB and VB of the semiconductor. While the two-step method includes concomitant photoexcitation of two semiconductors, hydrogen (H+/H2) and oxygen (H2O/O2) evolution occurs individually at different semiconductors similar to the usual Z-scheme photosynthesis.


image file: d3ta01629c-f2.tif
Fig. 2 Schematic of the function and essential role of semiconductor photocatalyst in decreasing the Gibbs free energy for water splitting reaction. (a and b) photocatalytic water splitting using a single-component or combination of two semiconductor systems.

In 1972, Fujishima et al.57 primarily described the use of TiO2 electrodes as photocatalysts for water-splitting reactions, since then, numerous semiconductors such as ZnO, Sb2O3, WO3, and TiO2 have been utilized as photocalaysts.58,59 However, the utilization of semiconductor photocatalyst requires the use of noble-metal cocatalysts in combination to realize higher water splitting efficiency.60,61 Hence, researchers are relentlessly exploring non-noble metal catalysts for water splitting to make the process cost-effective. In this regard, 2D materials such as metal nitrides, carbides, and phosphides have been explored as proficient catalysts for water reduction reactions. As discussed earlier, photon absorption, electron–hole pair separation and water redox reactions at the catalyst active sites are three crucial stages in photocatalytic water splitting. 2D material band gap can be modified by varying the number of stacked sheets, which fulfills the first condition. Secondly, due to the atomic thickness of 2D material, photogenerated charge-carriers have to cover lesser distances to reach the surface energetic sites. Further, the high specific surface area of 2D materials provides abundant active sites and higher accessibility for H+ adsorption. Lately, Ida et al.62 demonstrated how solar flux density and charge carrier lifetime in 0D nanocrystals and 2D crystal affects the water-splitting process (Fig. 3). A nanocrystal requires four excited electrons to ensure complete water splitting. But, photogenerated charge carriers possess a very short average lifetime of 1 μs before recombination. Therefore, to harvest four electrons, the nanocrystal should absorb four photons within a shorter period of 1 μs time. The solar flux density in the Earth's atmosphere is estimated to be 2000 μmol s−1 m−2. Thus, 4 ms is essential for photons to get adsorbed by the 0D nanocrystal and it is tough to afford such high solar flux to achieve water splitting using 0D nanocrystal. 2D materials, on the other hand, could minimize recombination rate by providing shorter pathways for charge-carriers to reach the active sites, thereby low solar energy flux is paramount in the case of 2D materials to achieve water splitting.


image file: d3ta01629c-f3.tif
Fig. 3 Schematic illustration of the photocatalytic water redox reaction under solar light illumination for (a) 0D crystal and (b) 2D system. Replicated with the agreement.62 Copyright 2014, American Chemical Society.

2D transition metal thio(seleno) phosphates (MPX3, X = S, Se) nanosheets containing Mn, Fe, Co, Ni, Cu, etc. metals have been investigated for photocatalytic HER due to their tunable composition, electronic structure, and wide band gap ranging from 1.3 to 3.5 eV.63–65 The electronic structure and properties of the MPX3 systems are investigated by optical absorption, photoluminescence, and DFT simulations. Zhang et al.66 calculated the band structure of MPS3 (M = Fe, Mn, Ni, Cd, Zn) and MPSe3 (M = Fe, Mn) systems using the Heyd–Scuseria–Ernzerhof (HSE06) functional, which is presented in Fig. 4a–f. This study specifies that the band gap of single-layer MPX3 varies from 1.90 to 3.44 eV, which surpasses the thermodynamic potential of 1.23 eV essential for water splitting. In addition, Yang et al.67 calculated the dynamic stability of MnPX3 (X = S or Se) nanosheets with the finite displacement method, and the corresponding phonon dispersion curve is shown in Fig. 4h and g. Interestingly, no imaginary phonon modes are observed in the phonon spectra, confirming the dynamic stability and they can be isolated as free-standing nanosheets. As displayed in the right panel of Fig. 4h and g, phonon bands below 8 THz are assigned to Mn–S/Se interactions, while the high frequency region above 8 THz corresponds to P2X6 moiety vibrations. Secondly, the radius of Se is greater than that of S, hence, the bond length of Mn–Se is larger than that of Mn–S and phonon bands of MnSe3 are shifted towards low frequencies as compared to the MnPS3 system.


image file: d3ta01629c-f4.tif
Fig. 4 Band structure near the Fermi level for (a) FePS3, (b) MnPS3, (c) NiPS3, (d) CdPS3, (e) ZnPS3, and (f) FePSe3 single-layers. Reprinted with permission.66 Copyright 2018, Wiley publication. Phonon dispersion and the corresponding density of states for (g) MnPS3 and (h) MnPSe3 nanosheets. Reprinted with permission.67 Copyright 2020, The Royal Society of Chemistry.

Apart from the band gap, the valence band and conduction band edges must straddle with H+/H2 reduction and H2O/O2 oxidation potentials. Liu et al.68,69 using HSE06 functional and partial density of states, calculated the orientation of CBM and VBM edges relative to water redox potential (Fig. 5a). Further, Zhang et al.66 by taking into consideration of stability of monolayered MPX3 systems and variation in their redox potentials with changes in pH, calculated the band edges based on HSE06 functional studies (Fig. 5b). Among the studied materials, FePSe3 displayed the narrowest band gap, which was further confirmed by optical absorption spectra as well. However, at pH = 7, FePSe3 displayed a CBM edge lower than water reduction potential and hence it is not suitable for H+/H2 reduction. Therefore, a particular MPS3 system can be chosen for either HER or OER based on the alignment of VBM and CBM edges relative to water redox potential. In addition, MnPS3, FePS3, and NiPS3 displayed strong optical absorption in the visible range, which indicates their high capability to harvest visible light (Fig. 5c). These simulations have been experimentally verified by Du et al.70 as well.


image file: d3ta01629c-f5.tif
Fig. 5 (a) Band edge potential of mono-layered MPX3 compounds. The energy scale is represented with respect to the normal hydrogen electrode (right y-axis) and vacuum level (left y-axis) in eV for reference. The redox potential of water is incorporated for assessment. Reproduced with permission.68 Copyright 2001, Macmillan Publishers Ltd. (b) Location of VBM and CBM edge potentials of monolayer MPS3 and MPSe3 calculated using HSE06 functional. The redox potential of water at pH = 0 and pH = 7 is incorporated for comparison. (c) The optical absorption coefficient for MPS3 compounds. The region between the red and purple lines indicates the visible range. Reprinted with permission.66 Copyright 2018, Wiley publication. (d) Schematic shows that the exfoliated ZnPS3 shows higher photocatalytic HER activity as compared to the bulk due to more exposed edge sites.112

Strong visible light absorption characteristics and high carrier mobility (625.9 cm2 V−1 s −1 for MnPSe3) specify the MPX3 system's potential capability for H2 generation under solar light.65,66,70 In addition, members of the MPX3 family possess the aforementioned properties along with abundant phosphorous active sites for HER in mono or few-layer form; thus, it is realistic to assume that they can be good candidates for many catalysis reactions as non-precious metal catalysts. 2D CdPS3 showed high HER performance of 10[thin space (1/6-em)]880 μmol h−1 g−1 while MnPS3 and FePS3 displayed 21.2 and 141.9 μmol h−1 g−1, respectively, under solar light.65,71,72 One drawback of 2D TMPCs is their weak oxidizing ability of the photogenerated holes because of the misaligned valence band maximum (VBM) edge for H2O/O2 oxidation potential and hence shows low stability because of photocorrosion.64,73 Combination of 2D MPX3 with other Cs4W11O35, CdS, TiO2, and g-C3N4 nanostructures is reported to be beneficial for water splitting as latter systems possess strong hole oxidizing ability.

In this perspective, we provide an overview of various 2D TMPCs utilized for photocatalytic and electrocatalytic HER applications. The chemical vapor transport (CVT) method for bulk MPX3 preparation is mentioned briefly and micromechanical exfoliation and liquid-phase exfoliation approaches to prepare 2D MPX3 from bulk crystal are discussed. The DFT approach demonstrating the significance of proper alignment of VBM and conduction band minimum (CBM) edges of 2D MPX3 concerning water redox potential towards HER is discussed at length. To improve the catalytic activity of MPX3, phase engineering, band structure engineering, and electronic state modulation by doping strategies have been employed (Fig. 6). Generation of the heterojunction between 2D MPX3 with other Cs4W11O35, CdS, TiO2, and g-C3N4 nanostructures is reported to improve hole oxidizing ability. Further, photogenerated electrons are readily available for water reduction due to an in-built electric field gradient and therefore heterostructures are likely to display a higher HER rate and good stability. Besides, ferroelectric CuInP2S6 displays good HER characteristics and its activity can be further enhanced by coupling with other photocatalytically HER-active materials such as g-C3N4 and ZnIn2S4. The permanent polarization electric field and large exciton binding energy characteristics of ferroelectric CuInP2S6 induce spatial separation of the photogenerated carriers and reduction potential of electrons, respectively, thereby showing good H2 evolution. Notably, the CuInP2S6/g-C3N4 heterostructure shows extraordinary stability under harsh photochemical conditions, suggesting its practical applicability.


image file: d3ta01629c-f6.tif
Fig. 6 Schematic highlights of 2D MPX3 compounds and their nanocomposites for photocatalytic and electrocatalytic HER applications.

2. Structure and composition

All the members in the MPX3 family show a typical characteristic that (P2S6)4− or (P2Se6)4− anion sublattice is present within each layered crystal. While the metal ions are dispersed around the (P2X6)4− bipyramids in a honeycomb arrangement.74 In Fig. 7b–d, we have shown the comparison of the crystal structure of the typical layered MoS2 and MPX3 systems. The van der Waals gap (the smallest gap between the S layers) in the MPX3 system containing first-row transition metals is in the range of 3.22–3.24 Å, which is relatively higher than that of metal dichalcogenide (MS2) systems.74 Various metal cations can be accommodated in the MPX3 system with a change in the size of the slab. As for the (P2X6)4− bipyramid structure modification, S2− remains intact while the P–P bond distance is altered to accommodate metal ions of different sizes. An increase in P–P bond distance of 2.148 to 2.222 Ao is reported on substituting Cd2+ with Ni2+ in the MPS3 system.75 Besides, the layer thickness of the MPX3 system is positively enhanced with an increase in P–P bond distance. On the cation side, the MPX3 system can accommodate either divalent metal ions or a combination of monovalent and trivalent metal ions. In literature, bimetallic Ni2P2S6−xSex,76 CuInP2S6−xSex77, and Sn2P2S6−xSex78 systems, with alloying of (P2S6)4− and (P2Se6)4− anion units are also reported. To date, MPX3 systems containing Fe, Co, and Ni metal ions have been widely explored due to magnetic ordering and their catalytic properties towards photocatalytic and electrocatalytic water splitting.
image file: d3ta01629c-f7.tif
Fig. 7 Composition and crystal structure of MPX3 compounds. (a) Periodic table highlighting the transition elements and their valence states constituting the MPX3 system. Crystal structure of (b) MoS2, (c) MIIPQ3 (MII = Zn, Cd, Mg; Q = S, Se) and (d) Ag0.5MIII0.5PQ3 (MIII = Sc, In; Q = S, Se).

2.1 Monometallic MPX3

If there is only one type of metal occupying the cationic site of MPX3, as shown in Fig. 7c, these systems are termed monometallic. It is convenient to study the MIIPX3 system by comparing it with MIVS2 (example: MoS2) crystal structure. MIIPX3 compounds can be visualized as a layered MS2 system with 1/3 of the M site occupied by P–P (P2) pairs, i.e., MII2/3(P)1/3S2. The sulfur atom comprises surfaces within the individual lamella. The octahedrally coordinated MIIPX3 system can be envisaged as MII ions occupying the 2/3 site and the rest 1/3 site is filled by P–P pairs. Further, the P–P pair bonded to six sulfur atoms to form an ethane-like (P2S6)2− sublattice, within individual phosphor atoms tetrahedrally coordinated to three sulfur atoms. Meanwhile, the sulfur atom coordinated with two MII ions and covalently attached to one phosphorous atom. Commonly, MIIPS3 systems exhibit the C2/m symmetry with a monoclinic crystal structure and show a layer stacking sequence of “AAA”.79 The monoclinic lattice parameter β value calculated for CoPS3 and FePS3 is 107.16°, indicating that these systems possess an undistorted monoclinic crystal structure. On the other hand, MIIPSe3 possess R3(−)H symmetry because of the higher P–Se bond distance and P–Se–P bond angle, whereas NiPSe3 shows C2/m symmetry similar to NiPS3.80,81

2.2 Bimetallic MPX3

The “M” atom of MPX3 can be exchanged with other metal atoms, containing home and hetero-charge substitutions. For instance, in systems such as Zn1−xFexPS3, Zn1−xNixPS3 (ref. 75) (0 ≤ x ≤ 1), and CdxFe1−xP2S6 (0 ≤ x ≤ 1)82 homo-charge distribution can be easily realized due to similar sizes of M1II and M2II ions. Most of the homo-charge mixed sulfides are monoclinic while selenides display trigonal symmetry. Heterochrage MPX3 containing ½ M1I (Ag or Cu) and ½ M2III trivalent metal ion (Cr, V, Al or In) where “x” being S or Se atom are also reported. These mixed-cationic sulphides broadly show two stoichiometric ratios; M1+M3+[P2S6]4− (e.g., AgInP2S6,83 AgScP2S6 and CuVP2S6 (ref. 84)) and M21+M2+[P2S6]4− (e.g., Ag2MgP2S6 and Ag2MnP2S6).79 Among these sulfides, AgInP2S6, and AgScP2S6 are trigonal and centrosymmetric, while CuVP2S6 is non-centrosymmetric. Mixed selenides exist in two stoichiometry: M1+M3+[P2Se6]4− (e.g., CuAlP2Se6 (ref. 85) with random Cu1+ and Al3+ arrangement) and M4/33+Y2/3[P2Se6]4−, where Y signifies an unfilled site (e.g., In4/3P2Se6).86

3. Synthetic methods

Production of single-layer graphene from graphite by the scotch-tape method has opened a plethora of opportunities for researchers to realize other atomically-thin 2D materials such as MoS2, phosphorene, and bismuthene.87,88 Lee et al. successfully adapted micromechanical exfoliation of bulk MnPS3 to prepare a few-layer MnPS3 of high-quality and large lateral size.89Fig. 8a–c shows the optical microscopy and atomic force microscopy (AFM) images of the few-layer MnPS3. Using the AFM height profile graph one can easily find out the number of stacked layers and lateral dimension of the nanosheet. For large-scale exfoliation, physical routes such as ball milling and liquid-phase exfoliation90,91 in solvents such as N, N-dimethylformamide, and N-methyl-2-pyrrolidone have been reported. In addition, the chemical vapor deposition (CVD) strategy is also used to prepare monolayers of MnPS3.65 The CVD strategy involves simultaneous phosphorization and sulfurization of the MnO2 template using sulfur and phosphor precursors, respectively in the two-zone furnace. MnPSe3 has also been synthesized by the CVD method by utilizing a Se source instead of S powder. Recently, Li et al.92 exfoliated bulk Ni2PS3 in an electrolyte containing tetra-n-butylammonium tetrafluoroborate as cationic intercalation salt by applying bias voltage (−3 V) to the bulk electrode and obtained exfoliated sheets in higher yields. The mechanism involves the intercalation of tetra-n-butylammonium salts, which considerably declines the van der Waals attractions between the stacked layers. Then, after applying voltage for a longer time, the intercalated tetra-n-butylammonium decomposes into the gaseous species which assists in delamination of the sheets from stratified bulk NiPS3 crystals.
image file: d3ta01629c-f8.tif
Fig. 8 (a) Optical microscopy, (b) AFM, and (c) conductive AFM image of few-layer MnPS3 obtained by the scotch-tape exfoliation method. Adapted with permission.89 Copyright 2016, AIP Publishing. (d) Schematic representation of the CVT method adapted for the preparation of MPX3 crystals and (e) photographs of the bulk NiPS3, FePS3, MnPS3, and FePSe3 prepared by the CVT method. TEM images of the exfoliated (f) Ag0.5In0.5PS3 and (g) Ag0.5In0.5PSe3 nanosheets.93

Rao and coworkers93 extended the chemical vapor transport (CVT) method to prepare monometallic (MnPS3, FePS3, NiPS3, ZnPS3, CdPS3, FePSe3, and CdPSe3) and bimetallic Ag0.5In0.5PS3 and Ag0.5In0.5PSe3) MPX3 systems (Fig. 8d and e). The synthesis involved heating the stoichiometric ratios of the constituent elements sealed in a quartz ampoule to ∼700 °C under an inert atmosphere. The pure bulk MPX3 compounds were formed at the other end of the quartz ampoule. The presence of sulfur and phosphor elements enhanced the vapor transport rate during material growth and well-developed crystals are obtained. The obtained bulk crystals were exfoliated in a water and ethanol mixture under sonication conditions. Fig. 8f and g shows the TEM images of the exfoliated Ag0.5In0.5PS3 and Ag0.5In0.5PSe3 systems, which revealed a few-layer nature. In literature, the majority of MPX3 at large-scale are prepared using the CVT method and this strategy is even applicable for doping other transition (Ni, Fe or Co) elements into MnPS3 (ref. 94) as well as for the synthesis of ferroelectric CuInP2S6 (ref. 95) and CuBiP2S6 materials.

4. Electrocatalysis

Electronic structure analysis can predict probable reasons behind the catalytic activity of various materials. Orbitals of both metal and non-metal impact the catalytic activity depending on the redox reaction and related interactions. Transition metals require dz2 orbitals for hydrogen adsorption, while oxygen evolution reactions require dxz and dyz orbital O2−, OH, and OOH adsorption. The catalytic system involving all three orbitals (dxz, dyz, and dz2) is considered to be superior for overall water splitting. In the case of a non-metallic system, P and S atoms are reported to be active sites for electrochemical reactions. The hydrogen adsorption free energy (ΔGH) of a catalyst is broadly recognized as a good descriptor in evaluating the H2 evolution activity of a particular material.47 The ideal catalyst should possess a thermoneutral H+ adsorption value. Song et al.96 calculated the ΔGH value on the basal plane and edges of NiPS3, FePS3, and NiFePS3 compounds using DFT calculations (Fig. 9a). These calculations suggested that the ΔGH value on the basal plane of NiPS3, FePS3, and NiFePS3 is very high (∼0.5 eV). Interestingly, 10% Fe-doped NiPS3 formed Ni0.9Fe0.1PS3; the ΔGH value on P and S edge sites was reduced to the thermoneutral value, indicating that the bimetallic MPX3 systems can be an ideal catalyst for HER. In another study, Sun et al.97 anchored different transition metals (e.g., Sc, Ti, V, Cr, Mn, Fe, Co, and Ni) on monolayer CuPS3 and examined the catalytic performance towards HER. As shown in Fig. 9b, V, Fe, and Ni decorated CuPS3 were situated at the peak of the volcano curve, suggesting that these systems are potentially good for water splitting. Further, the authors also calculated the influence of hydrogen cover rate and found that V anchored CuPS3 was catalytically active for HER over a wide hydrogen coverage. Apart from the ΔGH value, the electron transport ability at active sites of the catalyst plays a crucial role in determining the overall HER activity. In view of this, Mayorga et al.98 calculated the heterogeneous electron-transfer (HET) rates for CrPS4, MnPS3, FePS3, CoPS3, NiPS3, ZnPS3, CdPS3, GaPS4, SnPS3, and BiPS4 using cyclic voltammograms recorded in [Fe(CN)6]3/4 redox couple by the Nicholson's method (Fig. 9c). The HET performance of BiPS4 (koobs ∼0.024 cm s−1) was comparable with a glassy carbon substrate, which then decreased in the following order CdPS3 > CrPS4 > SnPS3 > CoPS. The HET rate is generally related to the inherent orbital orientation of the particular material and higher HET in the case of BiPS4 suggests that this catalyst is electrochemically more active than the other studied MPX3 systems. Further, the charge-transfer resistance (Rct) calculated for MPX3 using electron impedance spectroscopy is in the range of 0.5 to 6.5 kΩ with CdPS3 and BiPS4 showing lower Rct value, indicating that these MPX3 systems are semiconducting in nature (Fig. 9d). Incorporation of conducting matrix such as graphene and carbon nanotubes is reported to enhance charge-transfer characteristics and HER yield of some MPX3 systems.47
image file: d3ta01629c-f9.tif
Fig. 9 (a) Structure and the corresponding free energy diagram of hydrogen evolution at equilibrium for NiPS3, Ni0.9Fe0.1PS3, and FePS3. Reprinted with permission.96 Copyright 2019, American Chemical Society. (b) The Volcano curve of exchange current density for various transition metal anchored CuPS3. Reprinted with permission.97 Copyright 2022, Springer. (c and d) Cyclic voltammograms recorded in the ferri/ferrocyanide redox probe and the Nyquist plots of various 2D TMPCs. Reprinted with permission.98 Copyright 2017, American Chemical Society.

During electrolytic water splitting, the catalyst material is coated on the substrate (glassy carbon and indium tin oxide) and utilized as the anode. As depicted in Fig. 10a, the water-splitting electrolyzer consists of three main constituents, the anode, cathode, and electrolyte. Irrespective of the electrolytic media (acidic or basic), the water splitting reaction involves a change in free energy of 237 kJ mol−1 at 1 atm and 25 °C, corresponding to a theoretical cell voltage of 1.23 V to initiate water splitting.99 However, experimentally a voltage higher than 1.23 V is required due to mass, electrolyte, and transport resistances and also because of the sluggish kinetics of HER and OER reactions.100 The overpotential can be reduced by coating the catalyst at the anode and cathode surfaces, which decreases the energy barrier. Water splitting via HER involves the reduction of a H+ ion or H2O, based on the pH of the electrolytic medium.100,101

2H+ + 2e → H2 (acidic)

2H2O + 2e → 2OH + H2 (basic)


image file: d3ta01629c-f10.tif
Fig. 10 Schematic of the (a) water-splitting electrolyzer and (b) possible reaction routes for electrocatalytic HER in the acidic medium. Reproduced with permission.122 Copyright 2019, Elsevier. (c and d) Linear sweep voltammogram (LSV) curves obtained in 0.5 M H2SO4 and (e) Tafel plots for various bulk MPX3. Reproduced with permission.123 Copyright 2017, American Chemical Society. (f) Free energy diagram of the H2 evolution for NiPS3 and LSV curves of bulk NiPS3, exfoliated NiPS3, and NiPS3/RGO composites. Reproduced with permission.106 Copyright 2016, Wiley publication.

Under acidic conditions, HER follows three possible reaction pathways, namely Volmer, Tafel, or Heyrovsky pathways. The Volmer step involves the reaction of electrons with H3O+ to yield catalytically adsorbed H+ ion (Had) (2H+ + e → Hads). Subsequently, H2 evolution occurs either by following the Tafel (2Hads → H2) or Heyrovsky (Hads + H++e → H2) pathway or both (Fig. 10b).41–43 Irrespective of the paths by which H2 evolution happens, Had is always associated with the water reduction reaction. The Tafel plots could help in obtaining the reaction kinetics and mechanistic pathways for any catalyst, which shows water splitting via HER. The slope corresponds to the log(current density) versus overpotential plot, in the cathodic potential (η) range, which gives the Tafel slope value. The Tafel slope value signifies the amount of additional potential essential to increase a current density by 10 mA cm−2. For an ideal catalyst Pt, the Tafel slope values of 30, 40, and 120 mV dec−1 indicate that the Volmer, Heyrovsky, and Tafel steps are the rate-determining steps.102,103 The charge carrier mobility and the number of active sites also play an important role in describing the intrinsic HER activity for MPX3. For instance, MnPSe3 shows high charge carrier mobility (electron and hole mobilities of 625.9 and 34.7 cm2 V−1 S−1, respectively), and the massive divergence in the carrier mobility values suggests that efficient charge-separation for H2 and O2 evolution reactions.66 Further, exfoliated layers show high water redox activity compared to the bulk due to increasing in-plane conductivity and more exposed edge sites.

2D MPX3-based electrocatalysts display H2 evolution activity starting from very low to high, some of them even showing activity comparable to benchmark Pt/C catalysts. Experimentally, the electrocatalytic HER performance is accessed by the onset potential with respect to a standard hydrogen electrode, Tafel slope, and charge-transfer resistance concerning the benchmark Pt/C catalyst. The lower the value of these factors, the superior the catalytic activity. We have compared these parameters of 2D MPX3-based catalysts in Table 1, which suggest that the MPX3 containing Ni, Co, and Fe atoms, bimetallic TMPCs and their composite with other 2D systems display superior activity. Generally, it has been described that bimetallic phosphochalcogenides show better activity than monometallic systems due to synergistic interactions. Some of the widely examined 2D MPX3 and their heterostructures are discussed in the following section.

Table 1 Comparison of electrocatalytic HER activity of 2D TMPCs and their nanocomposites reported in the literature
Catalysts Electrode substrate Overpotential at 10 mA cm−2vs. RHE (mV) Tafel slope (mV dec−1) Electrolyte Reference
a GCE – glassy carbon electrode.
Exfoliated MPX3 nanosheets
MnPS3 GCEa 835 ± 68 0.5 M H2SO4 124
MnPS3 GCE 632 272 0.5 M H2SO4 94
MnPS3 GCE 1090 ± 71 0.1 M KOH 124
MnPSe3 GCE 640 ± 87 0.5 M H2SO4 124
MnPSe3 GCE 992 ± 56 0.1 M KOH 124
FePS3 GCE 211 ± 3 42 0.5 M H2SO4 47
FePS3 GCE 337 ± 4 1M KOH 47
FePS3 GCE 673 ± 4 3.5 wt% NaCl 47
FePS3 GCE 530 56 0.5 M H2SO4 98
CoPS3 Graphite 222 ± 2 71 ± 5 0.5 M H2SO4 125
CoPS3 GCE 590 84 0.5 M H2SO4 98
NiPS3 GCE 297 69 0.5 M H2SO4 106
NiPS3 GCE 398 159 1M KOH 106
NiPS3 GCE 816 54 3.5 wt% NaCl 106
NiPS3 GCE 205 74 0.5 M H2SO4 126
[thin space (1/6-em)]
MPX3 nanosheet composite with reduced graphene oxide (RGO)
FePS3/RGO GCE 108 ± 2 54 0.5 M H2SO4 47
FePS3/RGO GCE 467 ± 3 1M KOH 47
FePS3/RGO GCE 192 ± 2 3.5 wt% NaCl 47
NiPS3/RGO GCE 178 55 0.5 M H2SO4 106
NiPS3/RGO GCE 281 48 1M KOH 106
NiPS3/RGO GCE 543 94 3.5 wt% NaCl 106
[thin space (1/6-em)]
Transition element doped MPX3 nanosheets
Ni0.95Fe0.05PS3 GCE 130 114 1M KOH 96
Ni0.9Fe0.1PS3 GCE 72 73 1M KOH 96
Ni0.85Fe0.15PS3 GCE 152 187 1M KOH 96
Ni0.95Mn0.05PS3 GCE 135 1M KOH 94
Ni0.90Mn0.10PS3 GCE 142 1M KOH 94
Ni0.85Mn0.15PS3 GCE 143 1M KOH 94
Ni0.97Co0.03PS3 GCE 112 103 1M KOH 49
Ni0.95Co0.05PS3 GCE 71 77 1M KOH 49
Ni0.93Co0.07PS3 GCE 105 110 1M KOH 49
Ni0.91Co0.09PS3 GCE 145 113 1M KOH 49
Co0.6(VMnNiZn)0.4PS3 GCE 66 65 1M KOH 107
MnFePS3 Carbon paper 102 49 1M KOH 127
NiCoFePS3 Ni foam 231 86 1M KOH 128
[thin space (1/6-em)]
Single-atom anchored MPX3 nanosheets
NiFePS3 GCE 356 170 1M KOH 109
CoFePS3 GCE 490 132 1M KOH 109
PdFePS3 GCE 471 128 1M KOH 109
[thin space (1/6-em)]
2D MPX3–based heterostructures
FePS3/MoS2 Ni foam 175 127 0.5 M H2SO4 50
FePS3/MoS2 Ni foam 168 107 1 M KOH 50
NiPS3/MoS2 GCE 112 64 1 M KOH 51
NiPS3/Ni2P GCE 85 82 1 M KOH 110
NiPS3/Ni foam Ni foam 74 86 1 M KOH 129
CoNiPS3/N-doped carbon GCE 140 60 1 M KOH 130


In recent work, Martinez et al.98 studied the electrocatalytic HER performance of a series of bulk MPX3 (M = Mn, Fe, Co, Ni, Zn, Cd, Ga, and Sn) in acid electrolytes. Fig. 10c and d show the HER polarization curves for these MPS3 compounds and the resulting onset potential values at 10 mA cm−2. Among these MPX3 compounds, only NiPS3 and CoPS3 showed lower overpotential, −590 and −530 mV (vs. RHE), respectively, than bare GCE (−890 mV vs. RHE), indicating that these systems can be beneficial for electrocatalysis.69,98 The high H2 evolution activity of NiPS3 and COPS3 is ascribed to preferential orientation of these systems in the (001) crystal plane, as these planes are reported to be highly HER active in the case of Ni2P catalyst.33,104 Further, the MPS3 system containing Ni and Co metals is estimated to display ΔGH* value near to thermoneutral potential, thereby display superior HER activity.49,92,105 In another study, Sampath and coworkers106 have shown that HER activity of bulk NiPS3 can be further improved by exfoliating bulk crystals into few-layers by liquid phase exfoliation. These few-layer NiPS3 showed stable H2 evolution over a long period in a wide pH range of 1 to 14, including in 3.5 wt% NaCl solution, which is close to seawater composition. Fig. 10f shows the iR-corrected LSV polarization curves for bulk and exfoliated NiPS3, which show the onset potential of −450 and −159 mV, respectively, in 0.5 M H2SO4, indicating a higher activity in the case of the latter. Further, bulk NiPS3 exhibited a Tafel slope value of −119 mV dec−1, while a few-layer sample revealed a value of −69 mV dec−1. Further, exfoliated NiPS3 showed an overpotential of 297, 398 and −816 mV in acidic (pH = 1), basic (pH = 14) and in neutral 3.5 wt% NaCl electrolyte, respectively, at 10 mA cm−2. These studies indicate that the kinetics of HER is more facile in the case of exfoliated samples due to an enhanced charge-transfer rate. The authors also showed by DFT calculation that the P atom is the most favorable site for hydrogen adsorption by evaluating the ΔGH* value (Fig. 10f).

Sampath and coworkers47 reported HER activity of the exfoliated FePS3 system as well. The reported overpotentials at 10 mA cm−2 for exfoliated FePS3 in pH = 1, 14, and in neutral 3.5 wt% NaCl were −211, −337, and −637 mV respectively. According to Sampath and coworkers, exfoliated NiPS3 showed higher overpotential at 10 mA cm−2 and hence lower activity than FePS3. On the contrary, Martinez et al.98 reported that bulk NiPS3 showed lower overpotential at 10 mA m−2 as compared to FePS3. Therefore, the physicochemical properties of these materials need to be investigated in detail to provide further insights into the activity and mechanism of H2 evolution. In addition, coupling of exfoliated NiPS3/FePS3 with conductive reduced graphene oxide (RGO) substrate further improved HER activity as reported by Sampath and coworkers.47,106 The onset potential values obtained for exfoliated NiPS3 and FePS3 were −159 mV and −95 mV, while NiPS3/RGO and FePS3/RGO show −62 and −50 mV, respectively. The charge-transfer resistance (Rct) value obtained from AC impedance measurements indicates that the Rct value for NiPS3 and FePS3 was lowered on coupling with graphene. This implied that the conductive RGO support improved the HER activity of the MPX3 system due to enhanced charge-transfer characteristics. Also, NiPS3/RGO and FePS3/RGO nanocomposites showed long-term stability, suggesting these materials can be highly efficient electrocatalysts for HER in a wide pH range.

As indicated earlier, 2D MPS3 showed marginal electrocatalytic performance due to poor electronic conductivity; thus conductive graphene is coupled to improve the HER performance. Even though, NiPS3/RGO and FePS3/RGO composite required overpotentials of −178 and −192 mV, respectively, to reach 10 mA cm−2 in a basic electrolyte. In general, the catalyst electrocatalytic activity can be improved either by increasing the active site through catalyst loading or exposing the edge site via nanostructuring, or enhancing the intrinsic activity of the individual site. With this in mind, Li et al.49 doped Co into NiPS3 and with Ni0.95Co0.05PS3 nanosheets achieved a current density of 10 mA cm−2 at an overpotential of as low as −48 mV (vs. RHE) with a Tafe slope of 77 mV dec−1 in 1.0 M KOH electrolyte (Fig. 11a). The synthetic procedure involved initially obtaining the bulk NiCoPS3 by heating the stoichiometric ratio of individual elements in a sealed quartz ampoule. Later, the bulk sample was ultrasonicated in N,N-dimethyl formamide solvent to obtain 2D nanosheets. The authors also prepared Ni1−xCoxPS3 containing various cobalt amounts (x = 0.03, 0.05, and 0.07) to know the effect of Co doping content on HER activity. The Rct values of 18.6, 12.1, 23.4, and 41.7 Ω were obtained for Ni1−xCoxPS3 with cobalt doping amounts of x = 0.03, 0.05, 0.07, respectively, while bare NiPS3 showed 82.3 Ω (Fig. 11b). The reduced Rct value in the case of Ni1−xCoxPS3 implied that Co doping improved intrinsic HER kinetics of NiPS3 through a higher charge-transfer rate. However, an increase or decrease of Co amount other than x = 0.05, more or fewer declines HER performance. Later, theoretical studies have shown that H affinity at the P site can be improved by Co doping along with electrical conductivity.48 In another study, Rakov et al.94 showed that Ni-doped (lower oxophilicity) MnPS3 displayed higher HER activity than the bare sample. With Ni doping, the ability of binding of the surface intermediate into the catalyst surface decreased, which in turn assisted the desorption steps such as the Tafel or Heyrovsky steps. In addition, doping of multi-metal atoms such as V, Mn, Ni, and Zn into CoPS3 increased the density of active sites due to the formation of high entropy alloys.107 The generated S edge sites and basal plane P sites improved H+ ion adsorption, whereas Mn ion boosted water dissociation during the Heyrovsky step. Recently, Megha et al.108 using DFT calculations showed that Sc, Y, and Mo elemental doping into the MnPSe3 lattice reduced the ΔGH value of pristine MnPSe3 to 1.41 eV, to 0.24, 0.18, and 0.23 eV, respectively. Further, the density of state (DOS) plots suggested that DOS at the Fermi energy was infinite for 25, 12.5, and 6.25% Sc elemental doping concentrations in MnPSe3, indicating activation of the inert basal plane of MnPSe3 towards HER (Fig. 11c and d). Apart from metal atom doping, non-metal carbon doping can transform semiconducting FePS3 into metallic form and show enhanced HER activity. The dopant with fewer electrons than the substitutional atom can improve electronic conductivity and also greatly activate surface sites for hydrogen adsorption. Further, C, N-codoping in FePS3 pushed ΔGH (−0.188 eV) value closer to zero and showed HER activity comparable to that of Pt/C with a small overpotential of 53.2 mV to achieve a current density of 10 mA cm−2.105 However, studies on non-metal doping in 2D MPX3 are very limited and these systems need to be investigated in detail both by computational and experimental studies.


image file: d3ta01629c-f11.tif
Fig. 11 (a) LSV polarization curves and (b) Nyquist plots for Ni1−xCoxPS3 containing various cobalt amounts (x = 0.03, 0.05, 0.07).49 (c and d) Density of states plots for 25 and 6.25% Sc element doped MnPSe3. Reproduced with permission.108 Copyright 2023, Elsevier. (e) Schematic illustration of anchoring single-atom catalysts (Pt, Co, and Ni) over 2D FePS3. (f) Gibbs free energy profile of FePS3 and Ni–FePS3. (g and h) LSV polarization curves for HER and the corresponding Tafel plots, (i) polarization curves for OER for FePS3, Ni–FePS3, Co–FePS3, and Pd–FePS3 catalysts. Reproduced with permission.109 Copyright 2022, Wiley publication.

Recently, Tang et al.109 showed that anchoring of single-atom (SA) catalysts such as Ni, Co, and Pd over 2D FePS3 can tune the surface electronic structure and reinforces water adsorption and dissociation capacity for both OER and HER. DFT calculations showed that Ni SA anchored FePS3 facilitated electron aggregation from Fe to Ni–S and enhanced electron-transfer rate at Ni and S sites (Fig. 11e and f). In addition, the electron-rich Ni site acted as an active site to diminish hybridization between O and Ni in Ni–FePS3, thereby, facilitating conversion from O* to OOH* group and reducing the barrier for OER. Ni–FePS3 delivered an outstanding performance as a bifunctional catalyst for water splitting with an overpotential of 356 and 287 mV at a current density of 10 mA cm−2 for HER and OER, respectively (Fig. 11g–i). This study indicates that engineering surface active sites at the atomic scale are an effective way to modulate the catalytic activity of 2D MPX3 systems.

MPX3 nanostructures with a large number of active phosphorus edges are considered to be efficient water-splitting catalysts by both theoretical and experimental studies. The generation of heterojunctions by coupling two or more different materials can incorporate the advantages of distinct materials into one hybrid. Huang et al.50 reported the 2D/2D FePS3/MoS2 heterojunction, in which interfacial coupling between FePS3 and MoS2 causes significant enhancement in H2 evolution performance. FePS3/MoS2 requires low overpotentials of only 175 and 168 mV to reach 10 mA cm−2 current density under acidic and basic conditions, respectively (Fig. 12a and b). Further, FePS3/MoS2 showed lower Tafel slope values of 107 and 125 mV dec−1 at pH = 1 and pH = 14, respectively, with little decay in potential even after 1000 cycles (Fig. 12c). The in situ grown MoS2 can provide more exposed active sites, which facilitate diffusion of active species and release of the formed H2 bubble. Secondly, the intercalated MoS2 layers can stop restacking of FePS3 layers, enhancing conductivity and exposing active sites of FePS3. Since the electron transfer resistance is higher in adjacent layers than along the MPX3 layer, reduction in the stacking of the number of FePS3 layers improves electronic conductivity and carrier flow to MoS2.50 In another study, Liu et al.51 synthesized NiPS3/MoS2 composite and demonstrated that internal polarization field (IPF) facilities hydrogen spillover (HSo) during HER from the MoS2 edge site to NiPS3. Secondly, IPF boosts the hydroxyl ion diffusion pathway, thereby enhancing the rate of oxygen evolution reaction (Fig. 12d). The NiPS3/MoS2 composite requires a low overpotential of 112 and 296 mV to reach 10 mA cm−2 current density for H2 and O2 evolution reactions, respectively. Further, the NiPS3/MoS2 composite delivered overall water splitting with a cell voltage of 1.64 V at 10 mA cm−2 and exhibited long-term stability up to 100 h (Fig. 12e). The chronopotentiometry study conducted for 50 h indicates that NiPS3/MoS2 does not lose any catalytic activity, implying the practical advantages of the catalyst in terms of stability (Fig. 12f). Besides, Liang et al.110 grew nickel phosphide nanostructures on NiPS3, which showed a lower bias voltage of 1.65 V for overall water splitting, even surpassing that of the commercial Pt/C//IrO2 electrocatalyst (Fig. 12g–i). Generation of the MPX3 heterostructure with MoS2 would be beneficial as a bifunctional catalyst since these systems decrease the energy barrier for hydrogen adsorption and also improve sluggish OER kinetics due to built-in electric field gradient.115


image file: d3ta01629c-f12.tif
Fig. 12 (a) Schematic of the in situ CVT method adapted for the growth of FePS3/MoS2 heterostructure. (b) Polarization curves for FePS3/MoS2 heterostructure. (c) Electrochemical durability of FePS3/MoS2 in 0.5 M H2SO4. Reproduced with permission.50 Copyright 2020, Elsevier. (d and e) Polarization curve of NiPS3/MoS2 as both anode and cathode catalyst for overall water splitting driven by a solar cell (∼1.65 V). (f) Chronopotentiometry study for NiPS3/MoS2 carried for 50 h. Reproduced with permission.51 Copyright 2012, Wiley publication. (g) Gibbs free energy of hydrogen adsorption estimated for NiPS3/Ni2P composite, NiPS3 (110), Ni2P (110) and Ni2P (110) planes. (h) Charge density distribution at the junction of NiPS3/Ni2P composite. (i) Polarization curve of the NiPS3/Ni2P heterocatalyst for overall water splitting compared with the benchmark Pt/C//IrO2 electrocatalyst. Reproduced with permission.110 Copyright 2019, American Chemical Society.

5. Photocatalytic water splitting

2D MPX3 compounds band gap ranges from 1.3 to 3.5 eV, which appears to be advantageous for optoelectronic and catalysis applications. Secondly, abundant active sites such as P and S edges of 2D MPX3 facilitate hydrogen adsorption and desorption, these characteristics extend their application as photocatalysts for water-splitting applications. In this context, Shifa et al.65 prepared MnPS3 and MnPSe3 nanosheets by the CVD method and examined their photocatalytic HER performance under solar light (AM 1.5G) without any sacrificial agent. Atomic force microscopy images revealed that MnPS3 and MnPSe3 show layer thicknesses of ∼6 and 28 nm with lateral dimensions of ∼1.5 and 0.45 μm, respectively (Fig. 13a and b). Tauc plots derived from UV-vis diffuse reflectance spectroscopy show a red shift in absorption edge for the MnPSe3 as compared to MnPS3, indicating the better light harvesting capability of the former system (Fig. 13c). As depicted in Fig. 13d, both MnPS3 and MnPSe3 showed good HER activity with H2 evolution rates of 3.1 and 6.5 μmol h−1 g−1, respectively. The high H2 evolution rate of MnPSe3 as compared to MnPS3 can be ascribed to the superior light-harvesting ability and high charge carrier mobility in the former system. Secondly, the higher electronegativity of Se compared to S of P2X64− cluster orbitals could result in better H+ adsorption and H2 desorption. Further, the addition of Na2S/Na2SO3 as a sacrificial electron donor enhanced the H2 evolution rate of the MnPS3 and MnPSe3 photocatalysts to 21.2 and 43.5 μmol h−1 g−1, respectively. The energy level of MnPSe3 as derived from the Mott–Schottky plots indicated that VB was just fractionally below the water oxidation potential, which defied the oxygen evolution characteristics (Fig. 13e and f). Even though MnPSe3 showed promising HER performance, most of the research on the MnPSe3 system still focused on studying fundamental physical characteristics.
image file: d3ta01629c-f13.tif
Fig. 13 AFM images of MnPS3 (a), MnPSe3 (b), Tauc plot (c), HER data (d), Mott–Schottky plots (e), and schematic illustration of alignment of VBM and CBM edge potentials with respect to water reduction potential derived from Mott–Schottky plots (f) for MnPS3 and MnPSe3. Reprinted with permission.65 Copyright 2018, Wiley publication. Two-dimensional NiPS3 HER data under a xenon lamp and in solar simulated light (g), ultraviolet photoelectron spectroscopy data (h), and schematic illustration of alignment of VBM and CBM edge potentials of NiPS3 with respect to water reduction potential derived from Mott–Schottky plots (i). Reproduced with permission. Copyright 2017, Elsevier.113

Layered metal thiophosphates containing Fe, Ni, Sn, and Zn have also been reported as efficient catalysts for photocatalytic hydrogen evolution reactions. For instance, Sendeku et al.111 synthesized Sn2P2S6 nanosheets with Pc monoclinic phase by chemical vapor conversion process employing SnS2, P, and S precursors (Fig. 14a and b) and further utilized for H2 generation from pure water under solar light irradiation. To reveal the light-harvesting ability of Sn2P2S6, the UV-vis diffuse reflectance spectrum was recorded, as shown in Fig. 14c. The absorption edge of Sn2P2S6 was up to 520 nm with a weak absorption in the 520–600 nm region. Meanwhile, the corresponding Tauc plot shown in the inset of Fig. 14c shows the estimated band gap (Eg) value of 2.25 eV. Further, Mott–Schottky plots gave a positive slope revealing the n-type feature along with a flat band potential of −0.48 V (Fig. 14d). Secondly, ultraviolet photoelectron spectroscopy (UPS) studies were also carried out to obtain the band alignment of Sn2P2S6 nanosheets. From the UPS spectrum, the VB edge of Sn2P2S6 was calculated to be at −5.68 versus the vacuum level, as shown in Fig. 14e. The CB level was calculated from the equation ECB = EVBEg, which was estimated to be at −3.43 eV versus the vacuum level. On the basis of the previous results, a schematic band diagram of Sn2P2S6 is described in Fig. 14e, which implied that the VB and CB edges were suitably oriented to employ a water redox reaction. Fig. 14f shows the time course hydrogen evolution curve of Sn2P2S6 under simulated solar light (AM 1.5G) without any sacrificial agent. The H2 evolution rate increases linearly over time with a rate of 202.06 μmol h−1 g−1, which is ∼10 times higher than previously discussed MnPS3 nanosheets. In another study, Zhao et al.112 prepared exfoliated ZnPS3 by a liquid-phase exfoliation from its bulk crystal and realized a high HER activity of 640 μmol h−1 g−1 under a xenon lamp using Na2S and Na2SO3 as sacrificial agents (Fig. 14g–i). While bulk ZnPS3 showed relatively low activity of 640 μmol h−1 g−1 under similar conditions. The enhanced activity in the case of exfoliated ZnPS3 is attributed to a high specific area and more exposed edge sites.


image file: d3ta01629c-f14.tif
Fig. 14 (a and b) Schematic of the synthesis of Sn2P2S6 nanosheets using a tubular furnace. UV-vis diffuse reflectance spectrum (c), Mott–Schottky plots (d), energy band diagram (e), and HER data (f) for Sn2P2S6 nanosheets. Reprinted with permission.111 Copyright 2021, The American Chemical Society. UV-vis extinction spectra (g), UPS spectra (h), and time course HER curves (i) for bulk and exfoliated ZnPS3 catalyst. Reprinted with permission.112 Copyright 2022, Elsevier.

In 2017, Wang et al.113 prepared NiPS3 nanosheets by a modified CVD method and studied their HER activity under xenon light without any cocatalyst or sacrificial agent. As shown in Fig. 7g, NiPS3 showed the HER activity of 26.42 μmol h−1 g−1 in neutral or pure water until the light irradiation was turned off. Further, the addition of the hole scavenger Na2S/Na2SO3 to the photosystem enhanced the H2 yield to 74.67 μmol h−1 g−1. While the NiPS3 catalyst under simulated solar light showed a lower HER activity of 6.46 μmol h−1 g−1 in the sacrificial agent-free water. The band structure of NiPS3 was studied by using ultraviolet photoelectron spectroscopy and electrochemical impedance measurements indicated that the VB of NiPS3 was not straddling with water oxidation potential and hence was not energetically favorable to offer holes for oxygen evolution (Fig. 7h and i). Therefore, some of the single-component MPX3 systems utilized for photochemical HER without the cocatalyst or sacrificial agent did not evolve oxygen due to misalignment of the VB relative to the water oxidation potential. Thus, the MPX3 catalyst also possesses the common disadvantage of the single-catalyst photosystem, wherein photoactivity decreases with time due to photocorrosion. The photogenerated electrons at CB are consumed by H+ to generate H2 while holes accumulate at the VB due to misalignment. These highly reactive holes further react with the semiconductor itself, which results in decomposition and reduced activity of the catalyst. Even with the presence of a sacrificial agent, the activity of ZnPS3 and NiPS3 nanosheets was reduced by more than half of the initial value only after 24 h of the hydrogen evolution test, which suggests deprivation in structural stability under photochemical conditions. Therefore, the activity and stability obtained with a single-component MPX3 system are still far from reality, and photocorrosion of these materials under photocatalytic conditions impedes their applicability as photocatalysts.

In an effort to increase the H2 evolution yield, Barua et al.93 examined the photocatalytic HER activity of a range of MPX3 systems (MnPS3, FePS3, NiPS3, ZnPS3, CdPS3, FePSe3, CdPSe3, Ag0.5In0.5PS3, and Ag0.5In0.5PSe3) using eosin Y as a photosensitizer and triethanolamine (TEOA) as a sacrificial agent. This study indicated that the combination of photosensitizer and sacrificial agent somewhat assisted in extending the structural stability and long-term H2 evolution activity. Experimentally, band gaps of all synthesized MPX3 systems were obtained using absorption spectra and the corresponding Taucs plots revealed band gaps varied in the range of 1.29–3.3 eV, suggesting that the band gap and band edges of MPX3 were thermodynamically suitable for the water reduction reaction. Fig. 15a shows the schematic of the probable mechanism of HER by MPX3 compounds in the presence of Eosin Y and TEOA. Initially, the dye molecule absorbs light to yield a photoexcited dye (EY*), which then transforms to a triplet excited state (EY3*) via an intersystem crossing. The intermediate EY3* then takes an electron from TEOA via reductive quenching forming EY species. The electrons from EY are then transported to the CBM of the MPX3 where H+/H2 reduction takes place.97 The HER activity of monometallic TMPCs varies in the order NiPS3 > FePS3 > CdPS3 ∼ MnPS3 ∼ ZnPS3 where NiPS3 displays the highest HER rate of 2600 μmol h−1 g−1 (Fig. 15b and c). Fig. 15d presents the variation in the H2 evolution rate of the monometallic MPX3 with a change in P–P bond distance. The enhancement in the P–P bond distance leads to a reduction in the population of electrons at the P center, which is the active site, and therefore overall H2 evolution rate decreases. Under similar conditions, bimetallic Ag0.5In0.5PS3 and Ag0.5In0.5PSe3 showed appreciable hydrogen yields of 1900 and 500 μmol h−1 g−1, respectively (Fig. 15c). The catalytic stability is the additional significant factor for an HER catalyst. Under photochemical conditions, the NiPS3 system showed steady H2 evolution up to 5 cycles (25 h) without any decrement in H2 evolution, suggesting long-term stability of the photocatalyst (Fig. 15e).


image file: d3ta01629c-f15.tif
Fig. 15 (a) Schematic illustration of the mechanism of HER by exfoliated TMPCs nanosheets under a xenon lamp using eosin Y photosensitizer and TEOA sacrificial agent. (b) The yield of hydrogen evolved using monometallic (MnPS3, FePS3, NiPS3, ZnPS3, CdPS3) phosphosulfides. (c) Comparison of HER activity of monometallic (MnPS3, FePS3, NiPS3, ZnPS3, CdPS3 and bimetallic (Ag0.5In0.5PS3, Ag0.5In0.5PSe3) phosphosulfides. (d) Variation of H2 evolution rate of monometallic phosphosulfides with p–p bond distance. (e) Cyclic stability curves for NiPS3 studied up to 5 cycles for 25 h.93

2D MPX3 compounds are coupled with other semiconductors possessing good oxidation capability of holes to enhance H2 evolution yield and avoid photocorrosion. Theoretically, Mi and coworkers114 investigated the electronic structure of the MnPSe3 heterostructure with MoS2 using DFT calculations, where different stacking patterns of mono-layer MnPSe3 with MoS2 were investigated, as shown in Fig. 16a and b. In some stacking patterns, spin splitting at VBM of MnPSe3 is evident due to the hybridization of the d orbital of Mn, which enhances electron mobility. Secondly, MnPSe3/MoS2 forms a type II heterostructure where the top of the VB is majorly contributed by MnPSe3 while the bottom of CB is from MoS2. Normally, the type I heterojunction possesses a symmetrical offset of potential barriers, where direct exciton transition occurs at the heterointerface. While in the case of type II heterojunction holes and electrons are accumulated on different sides of the heterointerface, leading to indirect exciton transition. In type II heterojunction, the spatial separation of photoexcited electrons and holes as they are localized in different sides of heterojunction, reduces the recombination rate and enhances the photocatalytic efficiency. In the case of MnPSe3/CrSiTe3 heterojunction type I heterostructure is formed due to a similar crystal structure and low lattice mismatch. However, under tensile strain the band alignment changes from type I to type II due to the transition from an indirect bandgap to a direct bandgap. The above studies demonstrated the possibility of the modulation of the electronic structure of MPX3 by forming heterojunctions, implying the potential applicability of heterojunctions for photocatalysis applications.


image file: d3ta01629c-f16.tif
Fig. 16 (a) Structure and side view of the charge density difference of the MnPSe3/MoS2 nanocomposite with various stacking orders V1–V5. (b) Total and partial DOS of V1, V2, and V4 arrangements. The Fermi level is represented by the vertical shadow line and set to zero. Reprinted with permission.114 Copyright 2017, Springer Nature.

Experimentally, Chen et al.73 anchored 0D Cs4W11O35 nanoparticles onto 2D MnPS3 and reported higher HER yield in the case of MnPS3–Cs4W11O35 as compared to bare MnPS3. 2D MnPS3 is reported to be a non-toxic direct band gap semiconductor with appropriate VBM and CBM band edge potentials for photocatalytic water redox reaction. However, higher electron–hole recombination rates and their inability to oxidize water to generate oxygen limit their photo applicability. Bare MnPS3 was reported to show a low H2 production yield of 21.2 μmol h−1 g−1 under solar light illumination. To prepare MnPS3–Cs4W11O35 composites, initially, MnPS3 sheets are exfoliated in NMP solvent and then tetrabutylammonium hydroxide-modified Cs4W11O35 nanoparticles are anchored onto the MnPS3 sheets by electrostatic interaction. The homogeneous distribution of Cs4W11O35 nanoparticles on the MnPS3 sheets via strong interfacial interaction of the S–W–O bond is evident in TEM images (Fig. 17a). The photocatalytic HER performance of the MnPS3–Cs4W11O35 hybrid was examined under solar light with the sacrificial Na2S/Na2SO3 solution. Fig. 17b depicts the photocatalytic H2 evolution rate of bare MnPS3, MnPS3–Cs4W11O35 hybrid composites with Cs4W11O35 mass percentages of 5.4, 10.9, and 16.5%. The H2 evolution rates for MnPS3–Cs4W11O35 with 5.4, 10.9 and 16.5% of Cs4W11O35 were 38, 99, and 58 μmol h−1 g−1, respectively, while bare MnPS3 showed HER activity of 21.2 μmol h−1 g−1only. With the addition of Cs4W11O35 nanoparticles, the photocatalytic HER activity increased first and then decreased. The maximum H2 evolution rate of 99.6 μmol h−1 g−1 was achieved with the MnPS3–Cs4W11O35 composite with a Cs4W11O35 mass percentage of 10.9%. In order to investigate the long-term stability of the MnPS3–Cs4W11O35 composite, photocatalytic HER experiments were performed for four cycles, and during each cycle, the quartz reactor is evacuated prior to the start of the new experimental cycle. Cyclic stability curves shown in Fig. 17d indicate that HER activity for the second cycle is significantly higher than the first one and there is stable H2 evolution after the second cycle. The stable H2 evolution rate after the second cycle to the fourth cycle signifies the long-term photostability of the MnPS3–Cs4W11O35 composite. Further, TEM images of the MnPS3–Cs4W11O35 composite after photocatalytic HER studies, shown in Fig. 17e, reveal that Cs4W11O35 nanoparticles are coupled with MnPS3 sheets even after 24 h reaction. The PXRD pattern of MnPS3–Cs4W11O35 recorded after HER studies also showed that the catalyst was stable (Fig. 17f). Based on UV-vis, photoluminescence, valence band XPS spectra, and electrochemical impedance spectra studies, Chen et al.60,73 proposed the formation of the Z-scheme heterojunction, which effectively suppresses the charge-hole recombination rate, charge-transfer ability, and hole oxidation ability (Fig. 17g). The internal polarization field generated due to the heterojunction formation generates more positive charges on MnPS3 and more negative charges on Cs4W11O35. Upon light irradiation, electrons and holes are produced in VB and CB of both MnPS3 and Cs4W11O35 components of the composite. Further, the internal electric field suppresses the electron transfer from CB of Cs4W11O35 to CB MnPS3 and facilitates the electron migration from CB of Cs4W11O35 to VB of MnPS3, which results in the formation of the Z-scheme heterojunction. Hence, photogenerated electrons left in the CB of MnPS3 are utilized for water reduction whereas holes in the VB of Cs4W11O35 are utilized for the oxidation of Na2S/Na2SO3.


image file: d3ta01629c-f17.tif
Fig. 17 (a) TEM image of the MnPS3–Cs4W11O35 heterojunction. (b and c) Time course hydrogen production data and photocurrent responses for MnPS3–Cs4W11O35 catalyst with 0, 5.4, 10.9, and 16.5% of Cs4W11O35, and these composites designated as MnPS3, 5.4MPSCWO, 10.9MPSCWO, and 16.5MPSCWO, respectively. (d) Cyclic stability curves for HER of MnPS3–Cs4W11O35 catalyst recorded for 4 cycles and 24 h. (e) TEM image of MnPS3–Cs4W11O35 after photocatalytic HER studies. (f) Raman spectra of MnPS3–Cs4W11O35 before and after photocatalytic HER studies and (g) schematic of the reaction mechanism of photocatalytic HER on Z-scheme MnPS3–Cs4W11O35 heterojunction. Adapted with permission.73 Copyright 2021, Elsevier. (h) Time course HER curves for NiPS3/CdS composite. (i) Cyclic stability curves for HER of MnPS3–Cs4W11O35 catalyst recorded for 4 h and (j) Schematic representation of generated type-I (straddling type) heterostructure between NiPS3 and CdS components. (k) XPS spectra of NiPS3/CdS composite with light on and off respectively. Adapted with permission.115 Copyright 2022, Springer Nature.

As indicated earlier, in the case of 2D NiPS3, due to rapid charge recombination rate and photo-corrosion issues because of misalignment of the VBM concerning water oxidation potential, their reported HER activity is marginal. To overcome the weak oxidizing ability of NiPS3, Ran et al.115 coupled 2D NiPS3 with 0D CdS photocatalyst, which can supply oxidative photogenerated holes rather than being utilized. Cadmium sulfide (CdS) was reported to show a narrow band gap of 2.4 eV at room temperature, which accounted for its proficient visible light absorption characteristics. Secondly, CdS possess suitable VB and CB positions for water redox reactions, which make it an excellent photocatalyst for water redox reactions. Ran et al. examined the photocatalytic water splitting performance of NiPS3/CdS under visible-light illumination (λ > 400 nm) using TEOA as a sacrificial agent. Under light illumination, NiPS3/CdS showed an extremely high H2 evolution rate of 13[thin space (1/6-em)]600 μmol h−1 g−1 while bare NiPS3 showed negligible activity under similar conditions (Fig. 17h). The stability of NiPS3/CdS for photocatalytic HER was examined for 9 h, with a cycle every 3 h. As shown in Fig. 17i, the H2 evolution rate at cycle 3 (6687 μmol h−1 g−1) accounted for 49.17% of that in the first cycle (13[thin space (1/6-em)]600 μmol h−1 g−1). This result implied the appreciable stability of the NiPS3/CdS catalyst. In addition, in situ XPS measurements were conducted to reveal the dissociation and migration pathways of photogenerated electrons and holes on the surface of the NiPS3/CdS catalyst (Fig. 17k). During the in situ XPS measurement, a light-emitting diode was utilized as the light source to excite the catalyst. As presented in Fig. 17k, under light irradiation Ni 2p, P 2p and S 2p signals shifted towards higher binding energy as compared to dark conditions, signifying that more photogenerated holes than electrons migrate to the surface of both NiPS3 and CdPS3 components. This study suggests that during photocatalysis, holes migrate to the surface and are captured by the sacrificial agent, TEOA, leaving the electrons for HER. To explain the origin of such high HER activity of the NiPS3/CdS catalyst, the authors proposed a formation of type-I (straddling type) heterostructure between NiPS3 and CdS based on photoluminescence, Mott–Schottky plots, ultrafast transient absorption spectroscopy, steady-state and transient-state surface photovoltage (SPV) spectroscopy studies (Fig. 17j). Upon light irradiation, electrons and holes are photogenerated in both NiPS3 and CdS systems. Due to strong interfacial electronic coupling between NiPS3 and CdS, the electrons and holes in CB and VB of CdS migrate to NiPS3. However, most of the holes generated in CdS are consumed by sacrificial hole quencher TEOA before migrating to NiPS3, resulting in the oxidation of TEOA. While a low fraction of photogenerated holes are transported from VB of CdS to VB of NiPS3. On the other hand, much more photogenerated electrons in the CB of NiPS3 are utilized for the water reduction reaction. The significantly high HER activity of NiPS3/CdS can be attributed to the decreased recombination rate of photogenerated electron–hole pairs and abundant active sites on NiPS3 for hydrogen evolution.

Xia et al.116 reported the formation of S-scheme heterojunction between TiO2 and 2D FePS3 by self-assembling TiO2 nanoparticles on FePS3 nanosheets. In this study, FePS3/TiO2 nanocomposites are obtained by adding TiO2 (P25, Degussa AG) nanoparticles to the exfoliated FePS3 suspension under grinding conditions at room temperature. FePS3/TiO2 showed a higher hydrogen evolution rate of 99.5 μmol h−1 g−1 using ethanol solvent as a sacrificial electron donor under UV light illumination (Fig. 18a). While bare 2D FePS3 showed negligible H2 evolution under similar conditions. The Mott–Schottky plot for the FePS3/TiO2 composite indicated that both positive and negative slopes were obtained in distinct regions of the composite, implying the successful formation of the p–n heterojunction between 2D FePS3 and TiO2 (Fig. 18b). Further, in situ atomic force microscopy (AFM) combined with Kelvin probe force microscopy (AFM-KPFM) was utilized to investigate the electron transport pathway of the FePS3/TiO2 composite under light irradiation. As shown in Fig. 18d, aggregation of the TiO2 nanoparticles on the surface of FePS3 was evident in the AFM image of the FePS3/TiO2 composite. Fig. 18e and f show the corresponding KPFM images of FePS3/TiO2 in the dark and with 365 nm UV light irradiation, respectively. Accordingly, the surface potential plots of FePS3/TiO2 in dark and UV light irradiation are displayed in Fig. 18g. The surface potential across the line is enhanced under light illumination as compared to dark conditions. Particularly, the surface potential at A1 is elevated by 124 mV. This result indicated that the photogenerated holes accumulate at the surface of FePS3 and TiO2 components under light illumination. Upon light excitation, photoexcited electrons and holes are produced in both FePS3 and TiO2 components. Then, photogenerated electrons are retained in the CB of the FePS3 and holes preserved in VB of TiO2 to carry water reduction and oxidation, respectively, attributed to a strong in-built electric field gradient between the individual components. Meanwhile, photogenerated electrons in the VB of TiO2 recombine with the holes in the CB of FePS3. Due to the generation of the S-scheme heterojunction, the reduction ability of photogenerated electrons in FePS3 and the oxidation ability of holes in TiO2 are preserved (Fig. 18c). Secondly, a reduced recombination rate between the photogenerated electrons and holes was also simultaneously achieved. The photogenerated holes in VB of TiO2 were utilized for the oxidation of ethanol to produce the oxidized products.


image file: d3ta01629c-f18.tif
Fig. 18 (a) Photocatalytic HER activity of FePS3/TiO2 with various amounts of FePS3. (b) Mott–Schottky plots for FePS3/TiO2 acquired in 0.5 M Na2SO4 at 3000 MHz. (c) Schematic of the reaction mechanism of photocatalytic HER on S-scheme FePS3/TiO2 components. (d and e) Atomic force microscopy (AFM) and Kelvin probe force microscopy (AFM-KPFM) image of FePS3/TiO2 in dark conditions. (f) KPFM image with UV light irradiation. (g) The corresponding surface potential plot of FePS3/TiO2 in dark and UV light irradiation. Reprinted with permission.116 Copyright 2018, Wiley publication.

In recent years, 2D ferroelectric materials such as CuInP2S6, CuBiP2S6, and AgInP2Se6 are receiving increased attention as HER photocatalysts. The permanent polarization electric field in these materials serves as a driving force for the spatial separation of charges, thereby showing good HER activity.117–119 Secondly, large exciton binding energy suppresses the reduction potential of the photogenerated electrons, prompting electrons readily available for the water reduction reaction. Among these MPX3 compounds, CuInP2S6 experimentally showed good HER characteristics due to intermediate band gap and room temperature ferroelectricity.119 Lin et al.120 studied the photocatalytic HER performance of CuInP2S6, prepared by the solid-state reaction under visible light using a TEOA sacrificial electron donor. Bare CuInP2S6 nanosheets showed moderate HER activity of 18.3 μmol h−1 g−1 due to the high charge-transfer ability of the ferroelectric material. Further, the formation of 2D/2D heterojunction between CuInP2S6 and graphitic carbon nitride (g-C3N4) substantially accelerated the charge-transfer rate, due to which CuInP2S6/g-C3N4 displayed 2.5 times higher (45.1 μmol h−1 g−1) activity than that of bare CuInP2S6 (Fig. 19a–d). Transient photocurrent response studies indicated that the composite showed a higher photocurrent density of 2.92 μA cm−2, which was 1.62 and 2.25 times higher than that of individual CuInP2S6 and g-C3N4, respectively. This result implies the rapid charge separation and migration in the composite through high-speed microchannels. Further, photoluminescence studies implied a much lower emission intensity for the CuInP2S6/g-C3N4 composite compared to g-C3N4, indicating a much reduced electron–hole recombination rate in the composite. Based on the above results, the generation of type-II heterojunction between CuInP2S6 and g-C3N4 components is projected, as depicted in Fig. 19e. Upon light illumination, owing to the intrinsic dipole moment, CuInP2S6 can generate ferroelectric polarization fields with the deposition of negative and positive at opposite directions of the surface, which induces spatial charge separation. These photogenerated electrons at the CuInP2S6/g-C3N4 interface are utilized for water reduction to generate hydrogen, while holes are consumed by the sacrificial hole scavenger TEOA. Further, due to the in-built electric field gradient holes in the VB of g-C3N4 are transferred to the VB of CuInP2S6 by nanochannels, which further reduced the electron–hole recombination rate. Further, the cyclic stability curve shown in Fig. 19c implies that CuInP2S6/g-C3N4 still keeps high activity even after 5 cycles (15 h), suggesting their promising potential application in solar-energy conversion. In another study, Ren et al.121 synthesized a CuInP2S6/ZnIn2S4 heterostructure by a surfactant-assisted hydrothermal method and utilized it for photocatalytic H2 generation. As shown in Fig. 19f, bare ZnIn2S4 showed a quite inferior H2 evolution rate of 5.88 μmol h−1 g−1, while CuInP2S6/ZnIn2S4 displayed an enhanced activity of 76.2 μmol h−1 g−1. The band gap energy values for CuInP2S6 and ZnIn2S4 estimated using a UV-vis diffuse reflectance study were 2.37 and 2.02 eV, respectively (Fig. 19g). Owing to the matched band gap energy value between CuInP2S6 and ZnIn2S4, a type I heterostructure was effectively generated, which led to the efficient separation of the photogenerated electrons and holes (Fig. 19h). On visible-light irradiation, photogenerated electrons in the CB of CuInP2S6 migrated to the CB of ZnIn2S4 at the interface and were utilized for H+ reduction to yield H2. While photogenerated holes in the VB of CuInP2S6 migrated to the VB of ZnIn2S4 and were finally captured by the sacrificial agent TEOA. Markedly, due to the generation of the heterostructure, abundant charge-migration pathways were created, which reduced the charge recombination rate and improved the HER yield.


image file: d3ta01629c-f19.tif
Fig. 19 (a) Schematic of the photogenerated charge carriers in 2D CuInP2S6. (b) Time course H2 evolution data for g-C3N4(CN), CuInP2S6 (CIPS) and CuInP2S6/g-C3N4 (CN/CIPS). (c) Cyclic stability curve of HER for the CuInP2S6/g-C3N4 (CN/CIPS) heterojunction. (d) Temperature-dependent remnant polarization of CuInP2S6 inset: polarization-electric field (P-E) hysteresis loop obtained using piezoresponse force microscopy and (e) schematic representation of generated type-II heterostructure between CuInP2S6 and g-C3N4 components. Adapted with permission.120 Copyright 2020, Elsevier. (f) Time course H2 evolution data for ZnIn2S4 (ZIS) and CuInP2S6/ZnIn2S4 (CIPS/ZIS). (g) UV-vis diffuse reflectance spectrum of ZnIn2S4 (ZIS), CuInP2S6 (CIPS), and CuInP2S6/ZnIn2S4 (CIPS/ZIS) and (h) schematic illustration of generated type-I heterostructure at CuInP2S6/ZnIn2S4 interface. Adapted with permission.121 Copyright 2023, Elsevier. (i) Band edge potentials of CuInP2S6/Mn2P2S6 and CuInP2S6/Zn2P2Se6 ferroelectric heterostructures obtained using DFT+U method. The HER and OER potentials are shown for comparison. +P and −P designate CuInP2S6 in two oppositely polarized states.119

Theoretically, Huang et al.119 projected that the generation of ferroelectric hetero-interfaces such as CuInP2S6/Mn2P2S6 and CuInP2S6/Zn2P2Se6 would be beneficial for photocatalytic water splitting due to strong visible light absorption and type-II band alignment, which can facilitate rapid charge separation (Fig. 19i). To date, there are very limited experimental studies on ferroelectric hetero-interfaces containing 2D MPX3 compounds for photocatalysis and these systems need to be investigated in more detail. In Table 2, we have listed the photocatalytic HER activity of various 2D metal thiophosphates and their heterostructures reported in the literature along with the reaction conditions. It should be noted that, while comparing the HER activity between two individual systems, the strength of the light source (simulated solar light or xenon lamp, 300 W or 400 W), photosensitizer added and type of sacrificial agent (TEOA and Na2S/Na2SO3) need to be taken into consideration since these factors can affect the activity significantly. Secondly, different synthetic approaches generate MPX3 layers of various thicknesses, which further affect the HER activity as accessibility for catalytic edge (active) sites decreases with an increase in the layer thickness.

Table 2 Comparison of photocatalytic HER activity of 2D TMPCs and their nanocomposites with Cs4W11O35, CdS, TiO2, and g-C3N4
Photocatalyst system Synthesis method, layer thickness Light source Sacrificial agent HER activity (μmol h−1 g−1) Ref.
a CVD – chemical vapor deposition. b AM 1.5G – simulated solar light. c SS – solid state method. d LPE – liquid phase exfoliation.
MnPS3 CVDa, 6 nm AM 1.5Gb 3.1 55
MnPS3 CVD, 6 nm AM 1.5G Na2S/Na2SO3 21.2 55
MnPSe3 CVD, 28 nm AM 1.5G 6.5 55
MnPSe3 CVD, 28 nm AM 1.5G Na2S/Na2SO3 43.5 55
FePS3 CVD, 20 nm 300 W Xe lamp FeSO4 46.6 131
FePS3 CVD, 20 nm 300 W Xe lamp Triethanolamine 141.9 131
FePS3 CVD, 20 nm 300 W Xe lamp Na2S/Na2SO3 402.4 131
FePS3 SS methodc 300 W Xe lamp Triethanolamine 166.2 132
FePS3 LPEd,10 nm 300 W Xe lamp Triethanolamine 290.0 133
Porous FePS3 SS 7 nm 300 W Xe lamp Triethanolamine 305.6 132
NiPS3 CVD, 3.5 nm AM 1.5G 26.42 113
NiPS3 CVD, 3.5 nm AM 1.5G Na2S/Na2SO3 74.67 113
CuPS3 Hot injection 150W simulator Na2S/Na2SO3 2085 63
ZnPS3 LPE 300 W Xe lamp Na2S/Na2SO3 640.0 112
CdPS3 CVD AM 1.5G Na2S/Na2SO3 786.6 131
CdPS3 LPE, 3 nm 300 W Xe lamp Ethanol 10[thin space (1/6-em)]880.0 71
N-doped CdPS3 LPE 300 W Xe lamp DL-lactic acid 6280.0 134
InPS3 CVD AM 1.5G Na2S/Na2SO3 379.1 131
Sn2P2S6 CVD AM 1.5G Na2S/Na2SO3 202.06 135
ZnIn2S4 SS method Visible light (λ ≥ 420 nm) 5.88 121
[thin space (1/6-em)]
2D MPX3 HER activity reported with photosensitizer eosin Y (EY)
EY/MnPS3 LPE 400 W Xe lamp Triethanolamine 200.0 93
EY/FePS3 LPE 400 W Xe lamp Triethanolamine 600.0 93
EY/NiPS3 LPE, 4.9 nm 400 W Xe lamp Triethanolamine 2600.0 93
EY/ZnPS3 LPE 400 W Xe lamp Triethanolamine 100.0 93
EY/CdPS3 LPE 400 W Xe lamp Triethanolamine 300.0 93
EY/MnPSe3 LPE 400 W Xe lamp Triethanolamine 200.0 93
EY/FePSe3 LPE, 6.2 nm 400 W Xe lamp Triethanolamine 1700.0 93
EY/CdPSe3 LPE 400 W Xe lamp Triethanolamine 200.0 93
EY/Ag0.5In0.5PS3 LPE, 2.1 nm 400 W Xe lamp Triethanolamine 1900.0 93
EY/Ag0.5In0.5PS3 LPE, 3.1 nm 400 W Xe lamp Triethanolamine 500.0 93
[thin space (1/6-em)]
2D MPX3 heterojunctions
MnPS3/Cs4W11O35 300 W Xe lamp Na2S/Na2SO3 99.0 73
NiPS3/CdS 300 W Xe lamp Triethanolamine 13[thin space (1/6-em)]600.0 64
FePS3/TiO2 350 W Xe lamp Ethanol 99.5 116
MnPS3/carbon dot 300 W Xe lamp 339.63 136
CuInP2S6 300 W Xe lamp Triethanolamine 18.3 120
CuInP2S6/g-C3N4 300 W Xe lamp Triethanolamine 45.1 120
CuInP2S6/ZnIn2S4 SS method Visible light (λ ≥ 420 nm) 76.2 121


Among the reported various 2D MPX3 compounds, CdPS3 showed a significantly high H2 evolution rate of 10[thin space (1/6-em)]880 μmol h−1 g−1, while CuPS3 and ZnPS3 displayed 2085 and 640 μmol h−1 g−1, respectively (Table 1). However, in a single-component MPX3 system, the photogenerated holes are confined to the semiconductor and not available for water oxidation, hence resulting in photocorrosion of the semiconductor. The combination of sacrificial hole scavengers such as Na2S/Na2SO3 or triethanolamine with 2D MPX3 alleviates photocorrsion, and increases the activity to some extent. Generally, the formation of hetero-interfaces between 2D MPX3 compounds and material with good oxidation capability for holes such as Cs4W11O35 and CdS enhances the HER yield and photostability of 2D MnPS3. The combination of NiPS3 and FePS3 with CdS and TiO2 appears to be advantageous for photocatalytic HER. Mainly, the NiPS3/CdS heterostructure displayed a significantly high H2 evolution rate of 13[thin space (1/6-em)]600 μmol h−1 g−1. The generation of heterojunction effectively suppresses the electron–hole recombination rate and improves the charge-transfer and hole oxidation ability due to the internal polarization field. In Table 3, we have compared the photocatalytic HER activity of 2D MPX3 nanocomposites with other 2D materials (C3N4, metal sulfides, phosphides, and carbides) and their nanocomposites reported in the literature. The HER activity obtained with NiPS3/CdS surpasses the highest reported for TMDCs-based nanocomposites, indicating that 2D MPX3-based composites could be beneficial for large-scale hydrogen production with exceptional stability.

Table 3 Comparison of photocatalytic HER activity of mono and bimetallic 2D TMPCs and their nanocomposites with other 2D material naocomposites reported in the literature
2D nanomaterials
Photocatalyst system Light source Sacrificial agent HER activity (μmol h−1 g−1) Cyclic stability References
No. of cycles Total time of cycle (h) HER activity after cyclic stability as compared with 1st cycle
a TEOA – triethanolamine. b NR – not reported. c 2D phosphochalcogenides. d EY – eosin Y dye.
C3N4 300 W Xe lamp TEOAa 169.5 4 20 Negligible change 137
MoS2 300 W Xe lamp TEOA 339.5 NRb 138
ZnInS4 300 W Xe lamp Na2S/Na2SO3 3890.0 3 9 Negligible change 139
CdS 300 W Xe lamp MeOH 370.0 NR 140
FeSe2 300 W Xe lamp Na2S/Na2SO3 955.3 NR 141
Fe–BiOCl 300 W Xe lamp Na2SO4 3540.0 NR 142
Phosphorene 300 W Xe lamp Na2S/Na2SO3 512.0 2 20 Negligible change 143
CuPS3c 150W simulator Na2S/Na2SO3 2085.0 3 24 Negligible change 63
EY/NiPS3c,d 400 W Xe lamp TEOA 2600.0 5 25 Negligible change 93
CdPS3c 300 W Xe lamp Ethanol 10[thin space (1/6-em)]880.0 NR 71
[thin space (1/6-em)]
2D nanomaterial-based heterostructures
MoS2/CdS 300 W Xe lamp Lactic acid 4060.0 3 15 Negligible change 144
MoS2/TiO2 300 W Xe lamp TEOA 10[thin space (1/6-em)]046.0 3 12 Negligible change 145
MoS2/C3N4 300 W Xe lamp TEOA 1155.0 3 12 1098 μmol h−1 g−1 146
MoS2/C3N4 AM 1.5G TEOA 8300.0 6 16 Negligible change 147
Ti3C2/CdS 530 W Xenon lamp MeOH 1730.0 3 18 Negligible change 140
Cu3P/ZnInS4 300 W Xe lamp TEOA 5461.0 6 18 Negligible change 148
WS2/CdS 400 W Xe lamp Lactic acid 13[thin space (1/6-em)]132.0 4 12 Negligible change 149
FeSe2/C3N4 300 W Xe lamp Na2S/Na2SO3 1655.6 2 12 Negligible change 141
Ti3C2/TiO2 300 W Xe lamp MeOH 4672.0 5 25 Negligible change 150
C3N4/phosphorene 300 W Xe lamp TEOA 5850.0 3 12 Negligible change 151
MnPS3/Cs4W11O35c 300 W Xe lamp Na2S/Na2SO3 99.0 4 24 Negligible change 73
NiPS3/CdSc 300 W Xe lamp TEOA 13[thin space (1/6-em)]600.0 4 4 6800 μmol h−1 g−1 64


6. Conclusions and future opportunities

The global energy crisis and environmental pollution problems have become increasingly severe and the search for clean and sustainable energy resources is vital. Water splitting utilizing solar energy, electrical energy, or a combination of solar-electrical energy is demonstrated to be the environment-friendly route for future green hydrogen production. From the practical application viewpoint, designing a high-performance and cost-effective, yet metal-free catalyst, for photocatalytic/electrocatalytic water splitting is technologically essential and urgently needed.

In this perspective, we discussed the recent development of non-noble metal catalysts for water-splitting reactions with a focus on 2D MPX3 compounds by emphasizing novel strategies developed for activating MPX3 for photocatalytic and electrocatalytic HER. Theoretically, Gibbs free energy of hydrogen adsorption is considered to be a good descriptor in evaluating HER performance, and material with an approximate ΔGH value of zero is considered to be an ideal catalyst. DFT calculations revealed that pyrite type CoPS shows thermoneutral H+ adsorption, and also experimentally substantiated that CoPS possess Pt/C-like electrocatalytic H2 evolution activity. The exfoliation of bulk MPX3 into thin sheets enhances the HER activity due to the increased in-plane conductivity and more exposed edge sites. The electronic conductivity and HER activity of MPS3 sheets can be further improved by coupling a conducting matrix such as graphene. Doping of Co atoms into FePS3 and NiPS3 systems also enhances HER kinetics due to the electronic state modulation with doping. Coupling of TMPCs with other 2D systems such as MoS2 facilitates hydrogen spillover during H2 evolution from the MoS2 edge site to MPX3. These MPX3/MoS2 composites also showed good OER characteristics, indicating that these composites can be useful for overall water-splitting applications.

Strong visible light absorption characteristics, high carrier mobility, and thermoneutral H+ adsorption of 2D MPX3 make them suitable catalysts for photocatalytic water splitting. However, single-component MPX3 as a photocatalyst showed marginal HER activity and low stability due to misaligned VBM with respect to water oxidation potential. The development of MPX3 for photocatalytic HER is presently concentrating on enhancing the light absorption efficiency, the alleviation of photocorrosion, industrial scale H2 production, and improving stability. The addition of sacrificial hole scavengers such as Na2S/Na2SO3 or triethanolamine alleviates photocorrsion, and increases the activity and stability to some extent. Nevertheless, band gap engineering is required to enhance light absorption and reduce the photogenerated electron–hole recombination rate.

By coupling MnPS3 with a Cs4W11O35 semiconductor possessing good oxidation capability of holes, the HER activity of individual components can be enhanced. The generation of Z-scheme heterojunction between MnPS3 and Cs4W11O35 effectively suppresses the electron–hole recombination rate and improves charge-transfer and hole oxidation ability due to the internal polarization field. Further, a combination of NiPS3 and FePS3 with CdS and TiO2, respectively appears to be advantageous for photocatalytic HER. Particularly, the NiPS3/CdS heterojunction showed superior HER activity of 13[thin space (1/6-em)]600 μmol h−1 g−1, which is comparable to that of some of the highest reported transition metal-based nanocomposites reported in the literature.

In spite of the overwhelming progress in photocatalysis/electrocatalysis of 2D MPX3, there is still a lot of scope to enhance the performance and stability towards commercialization, some points are listed below.

6.1 Synthesis of atomically thin 2D MPX3 nanosheets

The chemical vapor transport method is the most widely utilized strategy to obtain bulk MPX3 and is then converted to a few-layer form by liquid-phase exfoliation. Most of the liquid-phase exfoliation strategies yield few-layer MPX3 nanosheets, which are semiconducting in nature. In the case of MoS2, lithium-intercalation of bulk material is exploited to produce atomically-thin metallic 2D sheets on a large scale. However, the lithium-intercalation approach in preparing atomically-thin MPX3 is not reported and these exfoliated can show enhanced HER characteristics if they are converted to the metallic form due to the increased in-plane conductivity and exposed edge sites. Further, the crystal structure and HER properties of these exfoliated systems need to be optimized with combined theoretical and experimental work to maximize their catalytic efficiency.

6.2 Surface functionalization

Similar to TMDCs, most of the 2D MPX3 systems show poor ambient stability due to reactive sulfur and phosphorous atoms. Covalent functionalization with electron-donating and withdrawing moieties can enhance the ambient stability and dispersibility of 2D MPX3, which is crucial for catalysis applications. Secondly, surface functionalization can improve the hydrogen adsorption/desorption ability of 2D MPX3 under various electrolytic conditions. The generation of heterojunction between functionalized 2D MPX3 with other 2D materials using covalent cross-linking or electro-restacking strategies could further improve charge-transfer characteristics and expose edge sites for HER.

6.3 Doping of non-metal atoms

Doping metal atoms of 2D MPX3 with other transition elements to form bimetallic compounds improves the HER characteristics through the electronic state modulation. However, a very limited reports on doping P and S/Se species, and these sites are reported to be active for hydrogen adsorption and desorption during HER. Therefore, controlled doping of non-metal elements, such as N and O can alter the electron donating/withdrawing ability of the active site, thereby improving HER characteristics. Secondly, co-doping of metal and non-metal elements would give rise to synergistic interactions. Theoretical insights of co-doping to modulate the electronic structure will assist experimental studies to tailor the transport properties of the system.

6.4 Ternary and mixed dimensional heterostructures

The generation of a ternary heterostructure of graphene, 2D MPX3, and layered double hydroxides (LDHs)/MXenes is worth exploring for overall water splitting as this system possesses optimum HER and OER sites with good electronic conductivity. It is important to note that some of the MPX3 systems, which show high HER activity suffer from issues related to stability and vice versa. In the case of a ternary superlattice-like structure, the 2D MPX3 layer is sandwiched between other 2D components, which can improve the structural and catalytic stability along with the activity. Further, coupling of 2D TMPCs with other 1D carbon nanotubes or with 1D MoS2 can generate a 3D porous framework with more accessibility for electrolytes, highly exposed edge sites, and interconnected electron-transfer channels, which will be useful for other HER-related catalysis applications.

After potential developments of proficient H2 evolution catalysts, recent research has demonstrated that 2D MPX3 catalysts can be highly efficient for photocatalytic and electrocatalytic water splitting. Solar water-splitting reactors need to be established to sustain large-scale H2 production at a low cost. In this context, photoelectrochemical cells (PECs) have the benefit of exploiting both solar and electrical energy and concurrently separating evolved H2 from O2. Band gap tunability and good electronic conductivity make 2D MPX3 a potential candidate for HER with high solar-to-hydrogen conversion efficiency. Also, strategies for safe H2 storage and transport need to be developed.

Conflicts of interest

The authors declare no conflicts of interest.

Acknowledgements

K. Pramoda gratefully acknowledges financial assistance from Science and Engineering Research Board, Government of India, Start-up Research Grant No. SRG/2022/000988, Vision Group on Science and Technology, Karnataka Science and Technology Promotion Society, Government of Karnataka (No. KSTePS/VGST/2021-22/CISEE/GRD-1010/66/2022-23/33) and the Minor Research Project Grant, Jain University (No. JU/MRP/CNMS/30/2023).

References

  1. R. Eisenberg, Science, 2009, 324, 44–45 CrossRef CAS PubMed.
  2. X. Tao, Y. Zhao, S. Wang, C. Li and R. Li, Chem. Soc. Rev., 2022, 51, 3561–3608 RSC.
  3. Y. Wang, H. Suzuki, J. Xie, O. Tomita, D. J. Martin, M. Higashi, D. Kong, R. Abe and J. Tang, Chem. Rev., 2018, 118, 5201–5241 CrossRef CAS.
  4. A. Kudo and Y. Miseki, Chem. Soc. Rev., 2009, 38 1, 253–278 RSC.
  5. I. Staffell, D. Scamman, A. Velazquez Abad, P. Balcombe, P. E. Dodds, P. Ekins, N. Shah and K. R. Ward, Energy Environ. Sci., 2019, 12, 463–491 RSC.
  6. D. A. Cullen, K. C. Neyerlin, R. K. Ahluwalia, R. Mukundan, K. L. More, R. L. Borup, A. Z. Weber, D. J. Myers and A. Kusoglu, Nat. Energy, 2021, 6, 462–474 CrossRef CAS.
  7. R. M. Ormerod, Chem. Soc. Rev., 2003, 32, 17–28 RSC.
  8. L. Fan, Z. Tu and S. H. Chan, Energy Rep., 2021, 7, 8421–8446 CrossRef.
  9. M. K. Singla, P. Nijhawan and A. S. Oberoi, Environ. Sci. Pollut. Res., 2021, 28, 15607–15626 CrossRef CAS PubMed.
  10. P. E. Dodds, I. Staffell, A. D. Hawkes, F. Li, P. Grünewald, W. McDowall and P. Ekins, Int. J. Hydrogen Energy, 2015, 40, 2065–2083 CrossRef CAS.
  11. A. Roy, M. Chhetri and C. N. R. Rao, in Advances in the Chemistry and Physics of Materials, 2019, pp. 376–398,  DOI:10.1142/9789811211331_0016.
  12. K. Pramoda and C. N. R. Rao, APL Mater., 2023, 11, 020901 CrossRef CAS.
  13. U. Maitra, U. Gupta, M. De, R. Datta, A. Govindaraj and C. N. R. Rao, Angew. Chem., Int. Ed. Engl., 2013, 52, 13057–13061 CrossRef CAS.
  14. Y. R. Girish, Udayabhanu, N. M. Byrappa, G. Alnaggar, A. Hezam, G. Nagaraju, K. Pramoda and K. Byrappa, J. Hazard. Mater. Adv., 2023, 9, 100230 CrossRef CAS.
  15. M. G. Walter, E. L. Warren, J. R. McKone, S. W. Boettcher, Q. Mi, E. A. Santori and N. S. Lewis, Chem. Rev., 2010, 110, 6446–6473 CrossRef CAS PubMed.
  16. K. Takanabe, ACS Catal., 2017, 7, 8006–8022 CrossRef CAS.
  17. Z. Li, M. Hu, P. Wang, J. Liu, J. Yao and C. Li, Coord. Chem. Rev., 2021, 439, 213953 CrossRef CAS.
  18. L. Sun, Q. Luo, Z. Dai and F. Ma, Coord. Chem. Rev., 2021, 444, 214049 CrossRef CAS.
  19. K. Pramoda, S. Servottam, M. Kaur and C. N. R. Rao, ACS Appl. Nano Mater., 2020, 3, 1792–1799 CrossRef CAS.
  20. N. Cheng, S. Stambula, D. Wang, M. N. Banis, J. Liu, A. Riese, B. Xiao, R. Li, T.-K. Sham, L.-M. Liu, G. A. Botton and X. Sun, Nat. Commun., 2016, 7, 13638 CrossRef CAS.
  21. K. Li, Y. Li, Y. Wang, J. Ge, C. Liu and W. Xing, Energy Environ. Sci., 2018, 11, 1232–1239 RSC.
  22. M. Chhetri, S. Sultan and C. N. R. Rao, Proc. Natl. Acad. Sci. U. S. A., 2017, 114, 8986–8990 CrossRef CAS PubMed.
  23. M. Chhetri, S. Maitra, H. Chakraborty, U. Waghmare and C. N. R. Rao, Energy Environ. Sci., 2016, 9, 95–101 RSC.
  24. K. Pramoda, U. Gupta, M. Chhetri, A. Bandyopadhyay, S. K. Pati and C. N. R. Rao, ACS Appl. Mater. Interfaces, 2017, 9, 10664–10672 CrossRef CAS.
  25. L. Yu, Q. Zhu, S. Song, B. McElhenny, D. Wang, C. Wu, Z. Qin, J. Bao, Y. Yu, S. Chen and Z. Ren, Nat. Commun., 2019, 10, 5106 CrossRef PubMed.
  26. D. C. Binwal, K. Pramoda, A. Zak, M. Kaur, P. Chithaiah and C. N. R. Rao, ACS Appl. Energy Mater., 2021, 4, 2339–2347 CrossRef CAS.
  27. U. Gupta and C. N. R. Rao, Nano Energy, 2017, 41, 49–65 CrossRef CAS.
  28. M. M. Ayyub, R. Singh and C. N. R. Rao, Sol. RRL, 2020, 4, 2000050 CrossRef CAS.
  29. Y. R. Girish, K. Jaiswal, P. Prakash and M. De, Catal. Sci. Technol., 2019, 9, 1201–1207 RSC.
  30. K. Pramoda, D. C. Binwal and C. N. R. Rao, Mater. Res. Bull., 2022, 149, 111697 CrossRef CAS.
  31. Y.-W. Cheng, J.-H. Dai, Y.-M. Zhang and Y. Song, J. Phys. Chem. C, 2018, 122, 28113–28122 CrossRef CAS.
  32. Y. Abghoui and E. Skúlason, J. Phys. Chem. C, 2017, 121, 24036–24045 CrossRef CAS.
  33. C. Hu, C. Lv, S. Liu, Y. Shi, J. Song, Z. Zhang, J. Cai and A. Watanabe, Catalysts, 2020, 10, 188 CrossRef CAS.
  34. X. Xu, H. Xu and D. Cheng, Nanoscale, 2019, 11, 20228–20237 RSC.
  35. C. Tsai, F. Abild-Pedersen and J. K. Nørskov, Nano Lett., 2014, 14, 1381–1387 CrossRef CAS.
  36. X. Tan, D. Zhao, Y. Sun, Z. Duan, X. Wang and X. Wu, CrystEngComm, 2022, 24, 6696–6704 RSC.
  37. X. Dai, K. Du, Z. Li, M. Liu, Y. Ma, H. Sun, X. Zhang and Y. Yang, ACS Appl. Mater. Interfaces, 2015, 7, 27242–27253 CrossRef CAS.
  38. D. Mosconi, P. Till, L. Calvillo, T. Kosmala, D. Garoli, D. Debellis, A. Martucci, S. Agnoli and G. Granozzi, Surfaces, 2019, 2, 531–545 CrossRef CAS.
  39. D. C. Binwal, M. Kaur, K. Pramoda and C. N. R. Rao, Bull. Mater. Sci., 2020, 43, 313 CrossRef CAS.
  40. B. Konkena, J. Masa, W. Xia, M. Muhler and W. Schuhmann, Nano Energy, 2016, 29, 46–53 CrossRef CAS.
  41. J. Wang, X. Yue, Y. Yang, S. Sirisomboonchai, P. Wang, X. Ma, A. Abudula and G. Guan, J. Alloys Compd., 2019, 819, 153346 CrossRef.
  42. C. C. Weng, J.-T. Ren and Z. Yuan, ChemSusChem, 2020, 13, 3357–3375 CrossRef CAS.
  43. L. Chen, J.-T. Ren and Z.-Y. Yuan, Green Chem., 2022, 24, 713–747 RSC.
  44. H. Zhang, T. Wei, Y. Qiu, S. Zhang, Q. Liu, G. Hu, J. Luo and X. Liu, Small, 2023, 19, 2207249 CrossRef CAS PubMed.
  45. J. Jia, W. Zhou, G. Li, L. Yang, Z. Wei, L. Cao, Y. Wu, K. Zhou and S. Chen, ACS Appl. Mater. Interfaces, 2017, 9, 8041–8046 CrossRef CAS PubMed.
  46. M. Cabán-Acevedo, M. L. Stone, J. R. Schmidt, J. G. Thomas, Q. Ding, H.-C. Chang, M.-L. Tsai, J.-H. He and S. Jin, Nat. Mater., 2015, 14, 1245–1251 CrossRef PubMed.
  47. D. Mukherjee, P. M. Austeria and S. Sampath, ACS Energy Lett., 2016, 1, 367–372 CrossRef CAS.
  48. S. Wang, B. Xiao, S. Shen, K. Song, Z. Lin, Z. Wang, Y. Chen and W. Zhong, Nanoscale, 2020, 12, 14459–14464 RSC.
  49. K. Li, D. Rakov, W. Zhang and P. Xu, Chem. Commun., 2017, 53, 8199–8202 RSC.
  50. H. Huang, J. Song, D. Yu, Y. Hao, Y. Wang and S. Peng, Appl. Surf. Sci., 2020, 525, 146623 CrossRef CAS.
  51. Y. Liu, Y. Chen, Y. Tian, T. Sakthivel, H. Liu, S. Guo, H. Zeng and Z. Dai, Adv. Mater., 2022, 34, 2203615 CrossRef CAS.
  52. J. Yang, D. Wang, H. Han and C. Li, Acc. Chem. Res., 2013, 46, 1900–1909 CrossRef CAS PubMed.
  53. T. Hisatomi, J. Kubota and K. Domen, Chem. Soc. Rev., 2014, 43, 7520–7535 RSC.
  54. M. A. Nadeem, M. A. Khan, A. A. Ziani and H. Idriss, Catalysts, 2021, 11, 60 CrossRef CAS.
  55. J. Ran, J. Zhang, J. Yu, M. Jaroniec and S. Z. Qiao, Chem. Soc. Rev., 2014, 43, 7787–7812 RSC.
  56. K. Pramoda and C. N. R. Rao, ChemNanoMat, 2022, 8, e202200153 CrossRef CAS.
  57. A. Fujishima and K. Honda, Nature, 1972, 238, 37–38 CrossRef CAS PubMed.
  58. K. Maeda, J. Photochem. Photobiol., C, 2011, 12, 237–268 CrossRef CAS.
  59. R. Abe, J. Photochem. Photobiol., C, 2010, 11, 179–209 CrossRef CAS.
  60. K. Wu, H. Zhu, Z. Liu, W. Rodríguez-Córdoba and T. Lian, J. Am. Chem. Soc., 2012, 134, 10337–10340 CrossRef CAS.
  61. H.-Y. Lin, H.-C. Yang and W.-L. Wang, Catal. Today, 2011, 174, 106–113 CrossRef CAS.
  62. S. Ida and T. Ishihara, J. Phys. Chem. Lett., 2014, 5, 2533–2542 CrossRef CAS PubMed.
  63. X. Zhang, K.-A. Min, W. Zheng, J. Hwang, B. Han and L. Y. S. Lee, Appl. Catal., B, 2020, 273, 118927 CrossRef CAS.
  64. J. Ran, H. Zhang, S. Fu, M. Jaroniec, J. Shan, B. Xia, Y. Qu, J. Qu, S. Chen, L. Song, J. M. Cairney, L. Jing and S.-Z. Qiao, Nat. Commun., 2022, 13, 4600 CrossRef CAS PubMed.
  65. T. A. Shifa, F. Wang, Z. Cheng, P. He, Y. Liu, C. Jiang, Z. Wang and J. He, Adv. Funct. Mater., 2018, 28, 1800548 CrossRef.
  66. X. Zhang, X. Zhao, D. Wu, Y. Jing and Z. Zhou, Adv. Sci., 2016, 3, 1600062 CrossRef PubMed.
  67. J. Yang, Y. Zhou, Q. Guo, Y. Dedkov and E. Voloshina, RSC Adv., 2020, 10, 851–864 RSC.
  68. J. Liu, X.-B. Li, D. Wang, W.-M. Lau, P. Peng and L.-M. Liu, J. Chem. Phys., 2014, 140, 054707 CrossRef.
  69. C.-F. Du, Q. Liang, R. Dangol, J. Zhao, H. Ren, S. Madhavi and Q. Yan, Nano-Micro Lett., 2018, 10, 67 CrossRef.
  70. K.-z. Du, X.-z. Wang, Y. Liu, P. Hu, M. I. B. Utama, C. K. Gan, Q. Xiong and C. Kloc, ACS Nano, 2016, 10, 1738–1743 CrossRef CAS PubMed.
  71. Y. Zhang, Y. Zhao, C. Bao, Y. Xiao, Y. Xiang, M. Song, W. Huang, L. Ma, H. Hou and X. Chen, J. Alloys Compd., 2022, 909, 164731 CrossRef CAS.
  72. T. A. Shifa, F. Wang, Z. Cheng, P. He, Y. Liu, C. Jiang, Z. Wang and J. He, Adv. Funct. Mater., 2018, 28, 1800548 CrossRef.
  73. K. Chen, J. Liu, Z. Huang, S. Zong, L. Liu and W. Tan, Int. J. Hydrogen Energy, 2021, 46, 33823–33834 CrossRef CAS.
  74. F. Wang, T. A. Shifa, P. Yu, P. He, Y. Liu, F. Wang, Z. Wang, X. Zhan, X. Lou, F. Xia and J. He, Adv. Funct. Mater., 2018, 28, 1802151 CrossRef.
  75. R. Brec, Solid State Ion., 1986, 22, 3–30 CrossRef CAS.
  76. G. Kliche, Z. Naturforsch. A, 1983, 38, 1133–1137 Search PubMed.
  77. J. Banys, J. Macutkevic, R. Grigalaitis and J. Vysochanskii, Solid State Ionics, 2008, 179, 79–81 CrossRef CAS.
  78. B. Zapeka, M. Kostyrko, I. Martynyuk-Lototska and R. Vlokh, Philos. Mag., 2015, 95, 382–393 CrossRef CAS.
  79. R. Samal, G. Sanyal, B. Chakraborty and C. S. Rout, J. Mater. Chem. A, 2021, 9, 2560–2591 RSC.
  80. A. Wiedenmann, J. Rossat-Mignod, A. Louisy, R. Brec and J. Rouxel, Solid State Commun., 1981, 40, 1067–1072 CrossRef CAS.
  81. S. Joergens and A. Mewis, Z. fur Anorg. Allg. Chem., 2004, 630, 51–57 CrossRef.
  82. S. Lee, P. Colombet, G. Ouvrard and R. Brec, Inorg. Chem., 1988, 27, 1291–1294 CrossRef CAS.
  83. Z. Ouili, A. Leblanc and P. Colombet, J. Solid State Chem., 1987, 66, 86–94 CrossRef CAS.
  84. G. Burr, E. Durand, M. Evain and R. Brec, J. Solid State Chem., 1993, 103, 514–518 CrossRef CAS.
  85. R. Pfeiff and R. Kniep, Z. fur Naturforsch. – B, 1993, 48, 1270–1274 CrossRef CAS.
  86. A. I. Brusilovets and T. I. Fedoruk, Zhurnal Neorganicheskoj Khimii, 1977, 22, 2085–2087 CAS.
  87. P. Vishnoi, K. Pramoda and C. N. R. Rao, ChemNanoMat, 2019, 5, 1062–1091 CrossRef CAS.
  88. K. Singh, A. Ohlan and S. Dhawan, Nanocompos.: New Trends Dev., 2012 DOI:10.5772/50408.
  89. S. Lee, K.-Y. Choi, S. Lee, B. H. Park and J.-G. Park, APL Mater., 2016, 4, 086108 CrossRef.
  90. W. Zhu, W. Gan, Z. Muhammad, C. Wang, C. Wu, H. Liu, D. Liu, K. Zhang, Q. He, H. Jiang, X. Zheng, Z. Sun, S. Chen and L. Song, Chem. Commun., 2018, 54, 4481–4484 RSC.
  91. K. Synnatschke, S. Shao, J. van Dinter, Y. J. Hofstetter, D. J. Kelly, S. Grieger, S. J. Haigh, Y. Vaynzof, W. Bensch and C. Backes, Chem. Mater., 2019, 31, 9127–9139 CrossRef CAS.
  92. X. Li, Y. Fang, J. Wang, B. Wei, K. Qi, H. Y. Hoh, Q. Hao, T. Sun, Z. Wang, Z. Yin, Y. Zhang, J. Lu, Q. Bao and C. Su, Small, 2019, 15, 1902427 CrossRef PubMed.
  93. M. Barua, M. M. Ayyub, P. Vishnoi, K. Pramoda and C. N. R. Rao, J. Mater. Chem. A, 2019, 7, 22500–22506 RSC.
  94. D. Rakov, Y. Li, S. Niu and P. Xu, J. Alloys Compd., 2018, 769, 532–538 CrossRef CAS.
  95. F. Liu, L. You, K. L. Seyler, X. Li, P. Yu, J. Lin, X. Wang, J. Zhou, H. Wang, H. He, S. T. Pantelides, W. Zhou, P. Sharma, X. Xu, P. M. Ajayan, J. Wang and Z. Liu, Nat. Commun., 2016, 7, 12357 CrossRef CAS PubMed.
  96. B. Song, K. Li, Y. Yin, T. Wu, L. Dang, M. Cabán-Acevedo, J. Han, T. Gao, X. Wang, Z. Zhang, J. R. Schmidt, P. Xu and S. Jin, ACS Catal., 2017, 7, 8549–8557 CrossRef CAS.
  97. Y. Sun, A. Huang, Z. Li, Y.-Q. Fu and Z. Wang, Electrocatalysis, 2022, 13, 494–501 CrossRef CAS.
  98. C. C. Mayorga-Martinez, Z. Sofer, D. Sedmidubský, Š. Huber, A. Y. S. Eng and M. Pumera, ACS Appl. Mater. Interfaces, 2017, 9, 12563–12573 CrossRef CAS.
  99. B. H. R. Suryanto, Y. Wang, R. K. Hocking, W. Adamson and C. Zhao, Nat. Commun., 2019, 10, 5599 CrossRef CAS PubMed.
  100. Z. P. Ifkovits, J. M. Evans, M. C. Meier, K. M. Papadantonakis and N. S. Lewis, Energy Environ. Sci., 2021, 14, 4740–4759 RSC.
  101. P. J. McHugh, A. D. Stergiou and M. D. Symes, Adv. Energy Mater., 2020, 10, 2002453 CrossRef CAS.
  102. C. N. R. Rao and M. Chhetri, Adv. Mater., 2019, 31, 1803668 CrossRef.
  103. M. M. Ayyub, M. Chhetri, U. Gupta, A. Roy and C. N. R. Rao, Chem. – Eur. J., 2018, 24, 18455–18462 CrossRef CAS.
  104. L. Feng, H. Vrubel, M. Bensimon and X. Hu, Phys. Chem. Chem. Phys., 2014, 16, 5917–5921 RSC.
  105. J. Wang, X. Li, B. Wei, R. Sun, W. Yu, H. Y. Hoh, H. Xu, J. Li, X. Ge, Z. Chen, C. Su and Z. Wang, Adv. Funct. Mater., 2020, 30, 1908708 CrossRef CAS.
  106. R. N. Jenjeti, M. P. Austeria and S. Sampath, ChemElectroChem, 2016, 3, 1392–1399 CrossRef CAS.
  107. R. Wang, J. Huang, X. Zhang, J. Han, Z. Zhang, T. Gao, L. Xu, S. Liu, P. Xu and B. Song, ACS Nano, 2022, 16, 3593–3603 CrossRef CAS.
  108. Megha and P. Sen, Int. J. Hydrogen Energy, 2023, 48, 21778–21787 CrossRef CAS.
  109. C. Tang, D. He, N. Zhang, X. Song, S. Jia, Z. Ke, J. Liu, J. Wang, C. Jiang, Z. Wang, X. Huang and X. Xiao, Energy Environ. Mater., 2022, 5, 899–905 CrossRef CAS.
  110. Q. Liang, L. Zhong, C. Du, Y. Luo, J. Zhao, Y. Zheng, J. Xu, J. Ma, C. Liu, S. Li and Q. Yan, ACS Nano, 2019, 13, 7975–7984 CrossRef CAS PubMed.
  111. M. G. Sendeku, F. Wang, Z. Cheng, P. Yu, N. Gao, X. Zhan, Z. Wang and J. He, ACS Appl. Mater. Interfaces, 2021, 13, 13392–13399 CrossRef CAS PubMed.
  112. Y. Zhao, C. Bao, Y. Zhang, K. Du, W. Huang and C. Su, Mater. Lett., 2022, 324, 132687 CrossRef CAS.
  113. F. Wang, T. A. Shifa, P. He, Z. Cheng, J. Chu, Y. Liu, Z. Wang, F. Wang, Y. Wen, L. Liang and J. He, Nano Energy, 2017, 40, 673–680 CrossRef CAS.
  114. Q. Pei, Y. Song, X. Wang, J. Zou and W. Mi, Sci. Rep., 2017, 7, 9504 CrossRef PubMed.
  115. J. Ran, H. Zhang, S. Fu, M. Jaroniec, J. Shan, B. Xia, Y. Qu, J. Qu, S. Chen, L. Song, J. M. Cairney, L. Jing and S.-Z. Qiao, Nat. Commun., 2022, 13, 4600 CrossRef CAS PubMed.
  116. B. Xia, B. He, J. Zhang, L. Li, Y. Zhang, J. Yu, J. Ran and S.-Z. Qiao, Adv. Energy Mater., 2022, 12, 2201449 CrossRef CAS.
  117. S. Kim, N. T. Nguyen and C. W. Bark, Appl. Sci., 2018, 8, 1526 CrossRef.
  118. Y. Li, J. Li, W. Yang and X. Wang, Nanoscale Horiz., 2020, 5, 1174–1187 RSC.
  119. S. Huang, Z. Shuai and D. Wang, J. Mater. Chem. A, 2021, 9, 2734–2741 RSC.
  120. B. Lin, A. Chaturvedi, J. Di, L. You, C. Lai, R. Duan, J. Zhou, B. Xu, Z. Chen, P. Song, J. Peng, B. Ma, H. Liu, P. Meng, G. Yang, H. Zhang, Z. Liu and F. Liu, Nano Energy, 2020, 76, 104972 CrossRef CAS.
  121. X. Ren, C. Tang, B. Xu, W. Tang, B. Lin and G. Yang, Mater. Lett., 2023, 333, 133654 CrossRef CAS.
  122. S. Gupta, M. K. Patel, A. Miotello and N. Patel, Adv. Funct. Mater., 2020, 30, 1906481 CrossRef CAS.
  123. C. C. Mayorga-Martinez, Z. Sofer, D. Sedmidubský, Š. Huber, A. Y. S. Eng and M. Pumera, ACS Appl. Mater. Interfaces, 2017, 9 14, 12563–12573 CrossRef.
  124. R. Gusmão, Z. Sofer and M. Pumera, Adv. Funct. Mater., 2019, 29, 1805975 CrossRef.
  125. N. Coleman, I. A. Liyanage, M. D. Lovander, J. leddy and E. G. gillan, Molecules, 2022, 27, 5053 CrossRef PubMed.
  126. J. Luxa, Š. Cintl, L. Spejchalová, J.-Y. Lin and Z. Sofer, ACS Appl. Energy Mater., 2020, 3, 11992–11999 CrossRef CAS.
  127. H. Zhang, G. Qi, W. Liu, S. Zhang, Q. Liu, J. Luo and X. Liu, Inorg. Chem. Front., 2023, 10, 2423–2429 RSC.
  128. M. Sial, H. Lin and X. Wang, Nanoscale, 2018, 10, 12975–12980 RSC.
  129. L. Fang, Y. Xie, P. Guo, J. Zhu, S. Xiao, S. Sun, W. Zi and H. Zhao, Sustainable Energy Fuels, 2021, 5, 2537–2544 RSC.
  130. Q. Liang, L. Zhong, C. Du, Y. Zheng, Y. Luo, J. Xu, S. Li and Q. Yan, Adv. Funct. Mater., 2018, 28, 1805075 CrossRef.
  131. Z. Cheng, M. Getaye and Q. Liu, Nanotechnology, 2019, 31 Search PubMed.
  132. J. Zhang, F. Feng, Y. Pu, X. Li, C. H. Lau and W. Huang, ChemSusChem, 2019, 12, 2651–2659 CrossRef CAS.
  133. Z. Cheng, T. A. Shifa, F. Wang, Y. Gao, P. He, K. Zhang, C. Jiang, Q. Liu and J. He, Adv. Mater., 2018, 30, 1707433 CrossRef PubMed.
  134. H. Li, N. Wells, B. Chong, B. Xu, J. Wei, B. Yang and Y. Guidong, Chem. Eng. Sci., 2020, 229, 116069 CrossRef.
  135. M. G. Sendeku, F. Wang, Z. Cheng, P. Yu, N. Gao, X. Zhan, Z. Wang and J. He, ACS Appl. Mater. Interfaces, 2021, 13(11), 13392–13399 CrossRef CAS PubMed.
  136. K. Chen, J. Liu, Z. Huang, S. Zong, L. Liu, W. Tan and Y. Fang, J. Colloid Interface Sci., 2022, 627, 438–448 CrossRef CAS.
  137. P. Yang, H. Ou, Y. Fang and X. Wang, Angew. Chem., Int. Ed., 2017, 56, 3992–3996 CrossRef CAS PubMed.
  138. L. Wang, Y. Hu, J. Xu, Z. Huang, H. Lao, X. Xu, J. Xu, H. Tang, R. Yuan, Z. Wang and Q. Liu, Int. J. Hydrogen Energy, 2023, 48, 16987–16999 CrossRef CAS.
  139. P. Hu, C. K. Ngaw, Y. Yuan, P. S. Bassi, S. C. Joachim Loo and T. T. Yang Tan, Nano Energy, 2016, 26, 577–585 CrossRef CAS.
  140. X. Chen, Y. Guo, R. Bian, Y. Ji, X. Wang, X. Zhang, H. Cui and J. Tian, J. Colloid Interface Sci., 2022, 613, 644–651 CrossRef CAS.
  141. J. Jia, W. Sun, Q. Zhang, X. Zhang, X. Hu, E. Liu and J. Fan, Appl. Catal., B, 2020, 261, 118249 CrossRef CAS.
  142. Y. Mi, L. Wen, Z. Wang, D. Cao, R. Xu, Y. Fang, Y. Zhou and Y. Lei, Nano Energy, 2016, 30, 109–117 CrossRef CAS.
  143. X. Zhu, T. Zhang, Z. Sun, H. Chen, J. Guan, X. Chen, H. Ji, P. Du and S. Yang, Adv. Mater., 2017, 29, 1605776 CrossRef.
  144. J. Xu and X. Cao, Chem. Eng. J., 2015, 260, 642–648 CrossRef CAS.
  145. B. Ma, P.-Y. Guan, Q.-Y. Li, M. Zhang and S.-Q. Zang, ACS Appl. Mater. Interfaces, 2016, 8, 26794–26800 CrossRef CAS.
  146. Y.-J. Yuan, Z. Shen, S. Wu, Y. Su, L. Pei, Z. Ji, M. Ding, W. Bai, Y. Chen, Z.-T. Yu and Z. Zou, Appl. Catal., B, 2019, 246, 120–128 CrossRef CAS.
  147. S. Masimukku, D.-L. Tsai, Y.-T. Lin, I. L. Chang and J.-J. Wu, Appl. Surf. Sci., 2023, 614, 156147 CrossRef CAS.
  148. R. Chen, H. Zhou, C. Han, P. Wang, R. Wang, Z. Liu, X. Ma and Y. Huang, ACS Appl. Energy Mater., 2022, 5, 12897–12906 CrossRef CAS.
  149. B. Archana, N. Kottam, S. Nayak, K. B. Chandrasekhar and M. B. Sreedhara, J. Phys. Chem. C, 2020, 124, 14485–14495 CrossRef CAS.
  150. C. Peng, T. Zhou, P. Wei, H. Ai, B. Zhou, H. Pan, W. Xu, J. Jia, K. Zhang, H. Wang and H. Yu, Chem. Eng. J., 2022, 439, 135685 CrossRef CAS.
  151. Y. Wang, C. Zeng, L. Wu, Y. Dong, Y. Zhang, D. Yang, W. Hu, J. Hao, H. Pan and R. Yang, J. Mater. Sci. Technol., 2023, 146, 113–120 CrossRef.

This journal is © The Royal Society of Chemistry 2023