Fabrication of amorphous subnanometric palladium nanostructures on metallic transition metal dichalcogenides for efficient hydrogen evolution reaction

Liang Mei a, Yuefeng Zhang a, Zimeng Ye a, Ting Han a, Honglu Hu a, Ruijie Yang a, Ting Ying a, Weikang Zheng a, Ruixin Yan a, Yue Zhang a, Zhenbin Wang a and Zhiyuan Zeng *ab
aDepartment of Materials Science and Engineering and State Key Laboratory of Marine Pollution, City University of Hong Kong, Hong Kong, People's Republic of China. E-mail: zhiyzeng@cityu.edu.hk
bShenzhen Research Institute, City University of Hong Kong, Shenzhen 518057, China

Received 9th March 2024 , Accepted 24th April 2024

First published on 2nd May 2024


Abstract

Fabricating solution-processable composite materials of transition metal dichalcogenides (TMDs) with ultrasmall noble metal structures employing an easy preparation method poses a significant challenge. In this study, we utilized a green, one-step synthetic method by directly employing electrochemical lithium intercalation-based exfoliated metallic TMD nanosheets (MoS2, WS2, and TiS2) to reduce palladium ions (Pd2+) to metallic Pd0, leading to the deposition on their surfaces. The resulting Pd nanoparticles (Pd NPs) in composites (Pd-MoS2, Pd-WS2, and Pd-TiS2) were found to be amorphous, with a size ranging from 0.81 to 1.37 nm. The impact of Pd NP size on hydrogen evolution reaction (HER) activity was elucidated. Among the fabricated composites, Pd-MoS2 exhibits the best HER performance, attributed to its smallest Pd NP size (0.81 nm). It shows an overpotential of 70 mV at a current density of 10 mA cm−2, along with a Tafel slope of 43 mV dec−1. These HER performance metrics surpass those of most Pd-decorated 2D catalysts.


1. Introduction

The electrocatalytic hydrogen evolution reaction (HER) is a crucial process in electrochemistry and renewable energy technology involved in the production of clean and sustainable hydrogen fuel.1–3 It involves the reduction of protons (H+) from solution to generate molecular hydrogen (H2) using an external electrical energy source. This reaction holds significant importance as hydrogen is considered a promising clean energy carrier due to its high energy density and environmentally friendly combustion, producing only water as a byproduct.4–6 The HER typically occurs at the cathode of an electrochemical cell, where an electrocatalyst facilitates the reaction by lowering the energy barrier required for the proton reduction. This catalyst enhances the kinetics of the reaction, making it more efficient and economically viable.7–9 The design and development of efficient electrocatalysts for the HER are critical for advancing hydrogen production technologies.10–12 Various materials, including noble metals as well as Earth-abundant and cost-effective catalysts have been explored to catalyze the HER.13–15 Researchers aim to improve the performance of electrocatalysts by enhancing their activity, stability, and cost-effectiveness. Tailoring the catalyst's surface structure, morphology, composition, and electronic properties is vital in optimizing its efficiency for hydrogen evolution.

Two-dimensional (2D) materials, such as graphene, transition metal dichalcogenides (TMDs), transition-metal carbides and/or nitrides (MXenes), have garnered significant attention for HER applications.14,16–18 These 2D materials possess atomically thin layers that provide high surface-to-volume ratios, exposing a significant number of active sites for catalysis.19 Their electronic band structures can be tailored by manipulating the number of layers, defects, or doping, allowing precise control over their catalytic activity.20–22 The fabrication of composites involving 2D materials and noble metals stands as one of the most effective strategies for the HER. At present, the methods used to prepare these composites often necessitate stringent conditions, such as toxic reducing agents,23 stabilizers,24 high temperature treatments,25,26 or complex procedures.27 In this study, we introduce a green preparation method that circumvents the need for using reducing agents or stabilizers and follows a one-step approach for preparing 2D TMDs (MoS2, WS2, and TiS2) decorated with amorphous subnanometric palladium nanostructures at room temperature. The metallic TMD nanosheets act as both a reducing agent and a stabilizer. Metallic TMD nanosheets, being electron-rich and possessing a matched redox potential, are capable of reducing Pd2+ to Pd0.28,29 The reduced Pd nanoparticles are stabilized on the TMD surface,30 ensuring their stability during the HER test. The resulting Pd-TMD composites demonstrate comparable HER performance to that of commercial Pt/C.

2. Experimental section

2.1 Materials

Molybdenum disulfide powder (MoS2, Innochem), tungsten disulfide powder (WS2, Macklin), titanium disulfide powder (TiS2, Sigma-Aldrich), N-methylpyrrolidone (NMP, Aladdin), polyvinylidene difluoride (PVDF, Sigma-Aldrich), natural graphite (Bay Carbon), copper foil (Aritech Chemazone Private), lithium foil (DodoChem), polypropylene (pp) film (Celgard 2300), lithium hexafluorophosphate dissolved in an ethylene carbonate, dimethyl carbonate and methylene carbonate mixture (1 M LiPF6 in EC[thin space (1/6-em)]:[thin space (1/6-em)]DMC[thin space (1/6-em)]:[thin space (1/6-em)]EMC, 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 vol%, DodoChem), platinum acetate (Pd(OAc)2, Macklin), and acetone (99.5%, Anaqua Global International Inc. Limited) were used. Milli-Q water (Millipore, Billerica, MA) was used in all experiments.

2.2 Synthesis of TMD (MoS2, WS2, and TiS2) nanosheet dispersions

2D TMD (MoS2, WS2, and TiS2) nanosheets were synthesized using a battery intercalation-based exfoliation method that we developed.31,32 In this process, lithium intercalation was carried out within a coin cell assembly, with bulk TMD-coated copper film serving as the cathode, lithium foil as the anode, and a 1 M lithium hexafluorophosphate (LiPF6) solution dissolved in a mixture of ethyl carbonate (EC), dimethyl carbonate (DMC), and ethyl methyl carbonate (EMC) in a 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 volume ratio as the electrolyte. The optimization of lithium intercalation in bulk TMDs was achieved by carefully controlling the cutoff voltage and current during the galvanostatic discharge process. Subsequently, the lithiated TMDs were extracted from the battery and subjected to sonication in deionized (DI) water, completing the exfoliation process. The resulting dispersions underwent multiple washes with DI water to eliminate any residual electrolyte and potential impurities. These TMD nanosheet solutions were employed as the basis for further characterization or subsequent integration with Pd nanostructures.

2.3 Synthesis of Pd-TMD (Pd-MoS2, Pd-WS2, and Pd-TiS2) composites

TMD nanosheets (MoS2, WS2, and TiS2) were individually dispersed in DI water at a concentration of approximately 40 mg L−1. To each TMD nanosheet solution, 100 μL of a 0.8 mg mL−1 palladium acetate solution was added. The solution was sonicated for 30 min to ensure thorough mixing and interaction. The resulting mixture underwent centrifugation at 9000 rpm for 5 min. Subsequently, the supernatant was carefully removed. The collected precipitate was re-dispersed in fresh DI water. This washing process was repeated three times to eliminate any residual reactants and impurities effectively. The purified precipitate containing Pd-TMD composites was re-dispersed in DI water for subsequent characterization and electrode preparation.

2.4 Material characterization

A transmission electron microscope (TEM, JEM 2100F) equipped with an energy-dispersive X-ray spectroscopy (EDS) detector was utilized for the examination of the morphology and elemental composition of the materials. The chemical state of the samples was characterized using an X-ray photoelectron spectrometer (XPS, Thermo Fisher ESCALAB XI XPS). The phase purity of the materials was confirmed by X-ray diffraction (XRD, D2 PHASER XE-T). Investigation of atomic vibration patterns was carried out using a Raman spectrometer (WITec alpha300 confocal Raman microscope, wavelength 532 nm). The surface charge of the TMD and composites was determined via zeta potential measurements (Malvern Zetasizer Nano series). These characterization techniques were employed to comprehensively analyze and understand the morphology, elemental composition, chemical state, phase purity, and atomic vibration patterns of the synthesized materials.

2.5 Electrocatalytic measurements

TMD and Pd-TMD composites were evaluated for their HER performance. For the catalyst ink preparation, each sample was mixed with ethanol and 5% Nafion perfluorinated resin solution (v/v = 25[thin space (1/6-em)]:[thin space (1/6-em)]1), forming the electrode coating materials. This mixture was subjected to a 30 min ultrasonication process to achieve optimal dispersion. Following this, 50 μl of the well-dispersed material was drop coated onto a 1.5 × 0.5 cm carbon paper electrode. The coated carbon paper electrode was air-dried at room temperature. The completely dried coated carbon paper electrode was then evaluated for its electrocatalytic properties using a typical three-electrode setup at room temperature. Linear sweep voltammetry (LSV) was performed in 0.5 M H2SO4 at a scan rate of 2 mV s−1. The setup included an Ag/AgCl electrode as the reference electrode, a graphite rod as the counter electrode, and carbon paper-coated electrocatalysts as the working electrode. All potentials were converted to values relative to a reversible hydrogen electrode, with the Ag/AgCl electrode consistently used as the reference in all measurements.

2.6 Computational methods

Spin-polarized density functional theory (DFT) calculations were carried out using the Vienna ab initio simulation package (VASP) code within the projected-augmented wave method.33,34 The exchange–correlation interactions were described by the generalized gradient approximation (GGA)35 within the Perdew–Burke–Ernzerhof (PBE) functional framework. The long-range van der Waals (vdW) interaction is treated with the empirical correction in Grimme's scheme (DFT-D3).36 The cutoff energy for plane-wave basis sets is 450 eV. The convergence thresholds for force and total energy are 0.05 eV Å−1 and 10−5 eV, respectively. The vacuum thickness was set to 16 Å to avoid the electronic interaction of periodic images. The Monkhorst–Pack scheme was adopted to sample the Brillouin region with a (3 × 3 × 1) k-point mesh grid for WS2, MoS2, and Pd(111), (2 × 2 × 1) for TiS2 and (4 × 3 × 1) for the Pd(211) supercell. The projected crystal orbital Hamilton population (pCOHP)37 analysis was employed to analyze the bonding/anti-bonding population of H–S and H–Pd. The Bader charge method was performed to analyze the changes in atomic charges.38 The Pd(111) slab and the Pd(211) slab contain four atomic layers; the atoms in the bottom two layers were fixed at a bulk truncated position, while the other atomic layers and the adsorbates were allowed to relax completely until meeting the convergence criteria.

The change in Gibbs free energy (ΔG) for each elementary step was calculated by exploiting the standard hydrogen electrode (SHE) model suggested by Nørskov et al.,39 given by:

ΔGH = ΔEH + ΔEZPETΔSH + ΔGpH

Here, ΔEZPE and ΔS represent the change in zero-point energies and entropy respectively, and T is the system temperature (298.15 K). ΔGpH = 2.303kBTpH represents the free energy contribution due to the variations in H concentration, where kB is the Boltzmann constant and the pH value was assumed to be zero for an acidic medium. ΔEH represents the energy required to increase the coverage by one hydrogen atom, which is calculated as follows:

ΔEH = E[base + nH] − E[base + (n − 1)H] − ½E[H2]

The base is the MoS2, WS2, TiS2, Pd(111), and Pd(211) catalyst. E[base + nH], E[base + (n − 1)H], and E[H2] are the total energy of the catalyst system with n and n − 1 adsorbed hydrogen atoms on the surface and of a gas phase H2 molecule, respectively.

3. Results and discussion

In this work, the metallic 2D MoS2, WS2, and TiS2 nanosheets were prepared from their bulk counterparts using an electrochemical lithium intercalation-based exfoliation method (refer to Experimental section and Fig. S1). The lithium intercalation process induces the phase transition because lithium intercalation involves the electron injection from the s orbitals of guest lithium to the d orbitals of the host transition metal atoms (Mo, W, and Ti), to maintain overall charge neutrality.40 Consequently, this method enables the fabrication of metallic phase MoS2, WS2, and TiS2 nanosheets with negative charge after exfoliation (Fig. S2–S4). We previously systematically optimized the electrochemical lithium intercalation conditions, including the discharging current and cutoff voltage, to ensure appropriate amounts of intercalated lithium in the interlayer of bulk TMDs for high-yield mono- or few-layer nanosheets upon exfoliation. Insufficient or excessive intercalated lithium ions can lead to the production of thick nanosheets or structural decomposition after exfoliation. For further details, see ref. 32. These metallic TMD nanosheets become enriched with additional electrons, capable of effectively reducing Pd2+ to Pd0 nanoparticles (Pd NPs) that are deposited on their surface (Fig. 1a). These fabricated composites (Pd-TMDs) were then employed to assess their electrocatalytic properties for the HER (Fig. 1b).
image file: d4qi00622d-f1.tif
Fig. 1 (a) Schematic illustration of the fabrication of Pd-TMD (Pd-MoS2, Pd-WS2, and Pd-TiS2) composites by the reduction of the Pd(OAc)2 precursor using the corresponding metallic TMD nanosheets as the reducing agent. (b) Pd-TMD composites used for the hydrogen evolution reaction.

After the exfoliation and washing steps, the TMD (MoS2, WS2, and TiS2) materials exhibit a nanosheet morphology characterized by a smooth surface and excellent crystallinity, as depicted in Fig. S5. Upon composite formation with Pd(OAc)2, TEM analyses of the Pd-modified TMD composites were conducted, and the results are presented in Fig. 2 and Fig. S6. Specifically, Fig. 2a, b, e, f, i and j emphasize the dense loading of monodisperse Pd nanoparticles (NPs) onto the MoS2, WS2, and TiS2 nanosheets, respectively. A detailed examination through high-resolution TEM images (Fig. 2c, g and k) revealed the amorphous structure of the Pd NPs. The measured d-spacing of 0.27 nm, shown in Fig. 2c and g, corresponds to the (100) plane of MoS2 and WS2, respectively.31 High-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) images of Pd-MoS2, Pd-WS2, and Pd-TiS2 are shown in Fig. 2d, h and l, respectively. The EDS mapping images directly validate the homogeneous dispersion of Pd NPs on MoS2, WS2, and TiS2 nanosheets (Fig. 2d, h, l and Fig. S7).


image file: d4qi00622d-f2.tif
Fig. 2 The morphologies of Pd-TMDs (Pd-MoS2, Pd-WS2, and Pd-TiS2). TEM images of (a and b) Pd-MoS2, (e and f) Pd-WS2, and (i and j) Pd-TiS2. The corresponding high-resolution TEM images are shown in (c) Pd-MoS2, (g) Pd-WS2, and (k) Pd-TiS2. EDX-mapping results are presented in (d) (Pd-MoS2, HAADF-STEM image and elemental mapping of Mo, S, and Pd), (h) (Pd-WS2, HAADF-STEM image and elemental mapping of W, S, and Pd), and (l) (Pd-TiS2, HAADF-STEM image and elemental mapping of Ti, S, and Pd).

The average sizes of Pd NPs in Pd-MoS2, Pd-WS2, and Pd-TiS2, as illustrated in Fig. 3a–c, were determined through statistical analysis to be 0.81 nm, 0.90 nm, and 1.37 nm, respectively. These findings confirm the successful fabrication of subnanometric Pd NPs on MoS2 and WS2 nanosheets. The XRD patterns exclusively exhibit the (002) peak for MoS2 and WS2, as well as the (001) peak for TiS2, respectively. Notably, no discernible peaks corresponding to Pd NPs were observed, further confirming the amorphous nature of the Pd NPs formed on these TMD nanosheets. These findings align consistently with the observations from the high-resolution TEM images depicted in Fig. 2c, g and k. The chemical states of pristine TMD nanosheets (MoS2, WS2, and TiS2) and those after Pd NP decoration (Pd-MoS2, Pd-WS2, and Pd-TiS2) were assessed through XPS measurements. All the Pd-TMD samples exhibit a strong Pd 3d signal in the XPS full spectra (Fig. S8), confirming the successful decoration of Pd NPs. As shown in Fig. 3e, Pd-MoS2, Pd-WS2, and Pd-TiS2 exhibit distinct doublets at approximately 335.4 and 340.7 eV, corresponding to Pd0 3d5/2–3/2 in the Pd 3d spectrum.41 This unequivocally indicates the reduction of Pd2+ to Pd0 within the system. Following the reaction, the Mo 3d5/2–3/2 spectrum (Fig. 3f), the W 4f7/2–5/2 spectrum (Fig. 3g), and the Ti 2p3/2–1/2 spectrum (Fig. 3h) within Pd-MoS2, Pd-WS2, and Pd-TiS2, respectively, exhibit discernible shifts toward higher binding energies when compared to their pristine TMD nanosheets. These observed peak shifts serve as direct evidence for electron transfer from the TMD nanosheets to Pd2+, facilitating the reduction of the latter to form Pd NPs. This charge transfer is also revealed through their Raman and zeta potential results. As shown in Fig. S4, after growing Pd NPs on TMD nanosheets, all the Raman peaks of Pd-TMDs shift to higher wavenumbers compared to the pristine TMDs. This shift in each peak is attributed to the charge transfer between the Pd NPs and TMDs. The peaks shift to higher frequencies, indicating direct charge transfer from TMDs to Pd NPs.41,42 Additionally, the zeta potential results of Pd-TMDs (Fig. S9) show that following the growth of Pd NPs, the zeta potential values of Pd-MoS2, Pd-WS2, and Pd-TiS2 increased to −1.57 mV, −3.07 mV, and −2.36 mV, respectively. This decrease in negative charge on the TMD surface further supports the observed charge transfer from TMDs to Pd nanoparticles after the reaction. Besides, the dominant metallic phase MoS2 and WS2, as well as a pure metallic phase in TiS2, are revealed through their XPS analysis.43–45


image file: d4qi00622d-f3.tif
Fig. 3 Pd NP size and chemical states of Pd-TMDs. The Pd NP size distribution analyses of (a) Pd-MoS2, (b) Pd-WS2, and (c) Pd-TiS2 were performed by counting over 100 Pd NPs using ImageJ software, and the data were fitted with the Gaussian function. (d) XRD patterns of Pd-TMDs (Pd-MoS2, Pd-WS2, and Pd-TiS2). (e) XPS Pd 3d spectra of Pd-MoS2, Pd-WS2, Pd-TiS2, and Pd(OAc)2. (f) XPS Mo 3d spectra of Pd-MoS2 and pristine MoS2. (g) XPS W 4f spectra of Pd-WS2 and pristine WS2. (h) XPS Ti 2p spectra of Pd-TiS2 and pristine TiS2.

The electrocatalytic performance of Pd-MoS2, Pd-WS2, and Pd-TiS2 composites was evaluated in a 0.5 M H2SO4 electrolyte using linear sweep voltammetry (LSV). As illustrated in Fig. 4a, d and g, in comparison with pristine MoS2, WS2 and TiS2 nanosheets, all Pd-decorated composites (Pd-MoS2, Pd-WS2, and Pd-TiS2) exhibited enhanced activity, as evidenced by the decreased overpotentials at a current density of 10 mA cm−2 in the polarization curves. This enhancement can be attributed to the synergistic coupling effect between TMD nanosheets and Pd NPs. From the Tafel plots, it is evident that after depositing the Pd NPs, the Tafel slope decreases from 101 mV dec−1 for MoS2 to 43 mV dec−1 for Pd-MoS2 (Fig. 4b). A similar enhancement in the catalytic activity is also observed for Pd-WS2 and Pd-TiS2 composites, with the Tafel slope lowering from 105 mV dec−1 for WS2 to 49 mV dec−1 for Pd-WS2 (Fig. 4e) and from 97 mV dec−1 for TiS2 to 61 mV dec−1 for Pd-TiS2 (Fig. 4h). Electrochemical impedance spectroscopy (EIS) was employed to assess the charge transfer resistance of Pd-TMDs and the pristine TMD electrodes. As demonstrated in the Nyquist plots with the equivalent circuit presented in Fig. 4c, f and i, the radii of the semicircles in the high-frequency region reflect the charge-transfer resistance, which signifies the resistance of charge transfer at the interface between the electrolyte and the electrode.46 This parameter is crucial for understanding the electrode kinetics under specific operating conditions. All Pd NP decorated composites (Pd-MoS2, Pd-WS2, and Pd-TiS2) exhibit smaller charge transfer resistance values compared to those of pristine MoS2, WS2, and TiS2. These further confirm the efficient charge transfer kinetics and agree with their polarization curves (Fig. 4a, d and g) and Tafel plots (Fig. 4b, e and h). Upon analyzing the various Pd NP sizes in Pd-MoS2, Pd-WS2, and Pd-TiS2, we observed a correlation between the particle size and the catalytic activity, revealing that smaller Pd NPs exhibit enhanced catalytic performance,47 as depicted in Fig. 4j. Cyclic stability is a crucial consideration for practical applications. Herein, all the Pd-TMD composites display long-term stability with no obvious decline in the current density at an overpotential of 130 mV (vs. RHE) for over 10 h (Fig. 4k). After the HER stability test, the peak positions in the XPS spectra of Pd 3d, Mo 3d, W 4f, and Ti 2p (Fig. S10), along with characteristic peaks in the Raman spectra (Fig. S11) of Pd-TMDs (Pd-MoS2, Pd-WS2, and Pd-TiS2), exhibit negligible differences after the durability test, indicating their structural stability post-testing. Furthermore, SEM (Fig. S12a–c) and HRTEM (Fig. S12d–f) images of Pd-MoS2, Pd-WS2, and Pd-TiS2 after the HER stability test reveal no obvious changes in morphology, with the monodisperse Pd nanoparticles remaining on the nanosheets without aggregation. Overall, the Pd-TMD catalysts fabricated in this study demonstrate excellent structural stability after the HER test. Besides, comparing the HER performance of Pd-TMDs with those of previously reported Pd-decorated 2D catalysts (Fig. 4l and ESI Table S1), it is evident that the Pd-MoS2 and Pd-WS2 composites developed in this study exhibit much better activity than most of the other catalysts, considering factors like the Tafel slope and overpotential at a current density of 10 mA cm−2. The superior performance can be attributed to the presence of subnanometric Pd NPs and the amorphous Pd structure in these composites.48 Consequently, they hold promise as alternatives to commercial Pt–C catalysts for HER applications.


image file: d4qi00622d-f4.tif
Fig. 4 HER performance of Pd-TMD composites. Polarization curves of Pd-TMD composites, pristine TMDs, and the commercial Pt/C catalyst are shown in (a) Pd-MoS2, (d) Pd-WS2, and (g) Pd-TiS2. The corresponding Tafel plots of (b) Pd-MoS2, (e) Pd-WS2, and (h) Pd-TiS2. EIS Nyquist plots of (c) Pd-MoS2, (f) Pd-WS2, and (i) Pd-TiS2 composites with the corresponding equivalent circuits shown in the insets. (j) The relationship of overpotentials (η) at 10 mA cm−2 and the Tafel slopes with the Pd NP size in Pd-MoS2, Pd-WS2, and Pd-TiS2. (k) Long-term stability test of Pd-TMDs under an overpotential of 130 mV (vs. RHE) in 0.5 M H2SO4. (l) Comparison of the overpotentials (η) at 10 mA cm−2 and the Tafel slopes between previously reported Pd decorated 2D catalysts and the Pd-MoS2, Pd-WS2, and Pd-TiS2 developed in this work. The details can be found in ESI Table S1.

To gain a more comprehensive understanding of the experimental phenomena, we use DFT calculations to elucidate the underlying reaction mechanism. The theoretical model of the Pd cluster, designed to be of equivalent size to the experimental system, is excessively large, resulting in substantial computational requirements in terms of time and cost. (111) and (211) are the low-index and high-index faces of Pd, and (111) has a lower surface energy (Fig. S13d), which is also consistent with the fact that the (111) face is generally considered to be the most stable face in the face-centered cubic structure (Pd). In addition, as the particle size increases, the structure begins to resemble more closely to the bulk Pd, predominantly exposing the representative (111) plane. Therefore, we choose the (211) and (111) planes to approximate the transition trend of particles from small (unstable surface) to large (stable surface).

Fig. S13a–c shows the structural models and possible adsorption sites of 1T-TMD (MoS2, WS2 and TiS2), Pd(111), and Pd(211) systems. TMD has 3 adsorption sites, and the ontop (T1) site is a stable adsorption site (Fig. S14). Pd(111) has 4 adsorption sites, with the most stable adsorption site being the FCC (F) site (Fig. S15a), and Pd(211) has 14 adsorption sites, with the most stable adsorption sites being H1 and H2 (Fig. S15b). Fig. 5a displays the thermodynamic Gibbs free energy diagrams of TMDs. H adsorption on MoS2 and WS2 exhibits significantly negative ΔG, suggesting that H has strong interactions with MoS2 and WS2 substrates, allowing the second H to absorb on the surface. The formation of the second H* is an uphill process with ΔG values of 0.91 eV and 0.52 eV for WS2 and MoS2, respectively, and is the rate-determining step (RDS) of the reaction. On TiS2, in contrast, the first H* formation is endothermic, which requires less energy compared to that of the RDS of WS2 and MoS2. Consequently, TiS2 exhibits better catalytic activity. The formation of H* is energetically supported on Pd(111) (Fig. 5b), resulting in a higher concentration of H atoms occupying the surface. The ΔG of H* with a coverage of 1 (9H*) is closest to 0, and it will be very difficult to adsorb an additional H atom (10H*), so the overpotential of Pd(111) is 0.21 V. Similarly, the first H* is easily formed on Pd(211) (Fig. 5c), and 9H* coverage is the RDS of the reaction with an extremely low ΔG value of 0.07 eV. In summary, the presence of Pd enhances the activity of HER and Pd(211) exhibits the highest catalytic performance, which provides a solid explanation for the improved catalytic activity observed in experiments using small-sized Pd particles. To gain a deeper understanding of the H adsorption mechanism, we present the relevant adsorption energy diagrams and electronic properties. Fig. 5e illustrates the charge density difference of H* at different coverages absorbed on various substrates. Larger electron clouds imply more charge transfer and thus exhibit strong interactions. MoS2 with one H* and WS2 with 2H* exist in the largest and smallest electron clouds, respectively, corresponding to the strongest and weakest interactions, which further elucidates the adsorption energy of H and the substrate (Fig. 5d). Quantitative relationships are also calculated via Bader charges (Fig. 5f). H bonded to Pd gains electrons from Pd, while H bonded to TMD loses electrons to S. The electron transfer between the second H* and S is minimal, implying their weak interaction, as shown in Fig. 5e. Furthermore, the integrated crystal orbital Hamilton population (ICOHP) analysis was also performed to quantify the population of electrons in bonding and antibonding molecular orbitals between H and S atoms and Pd atoms. As displayed in Fig. 5g, the ICOHP value for H–Pd in Pd(111) is more negative than that for H–Pd in Pd(211) (Fig. 5h), implying that H adsorption on Pd(111) is more energetically supported, which further explains the free energy diagram. Also, the trend of the bonding strength shown in Fig. S17–19 aligns with the energy. In conclusion, the introduction of Pd improves the catalytic activity compared to TMD, and smaller particle sizes demonstrate better catalytic performance; these calculated results well support the experimental data.


image file: d4qi00622d-f5.tif
Fig. 5 DFT calculations for clarifying the mechanism. Gibbs free energy diagram for the HER in (a) MoS2, WS2, and TiS2, (b) Pd(111), and (c) Pd (211). (d) The adsorption energy values of H* with a low average and a high coverage on different substrates and (e) the corresponding charge density difference diagrams. (f) Charge transfer of RDS H* on TMDs and different coverage H* on excellent Pd(211). (g) Crystal orbital Hamilton population (COHP) analysis of the interactions between H–Pd in Pd(111) and (h) Pd(211). Pd1, Pd2, and Pd3 are the three atoms bonded to H* on the Pd surface, as shown in Fig. S16.

4. Conclusions

In this study, amorphous subnanometric Pd NPs have been successfully fabricated on metallic TMD nanosheets (MoS2, WS2, and TiS2) through redox reactions between them. The metallic TMD nanosheets serve dual roles as both a reducing agent and a stabilizer, facilitating the reduction of Pd2+ to Pd0 and anchoring these ultrasmall Pd NPs on their surface using this solution-processable method. The resulting composites (Pd-MoS2, Pd-WS2, and Pd-TiS2) exhibited HER performance comparable to that of commercial Pt/C. This study introduces a novel and environmentally friendly synthetic strategy for fabricating cost-effective, easily producible amorphous ultrasmall noble metal-decorated TMD composites, contributing to the efficient generation of clean energy.

Author contributions

Liang Mei: conceptualization, methodology, investigation, data curation, writing – original draft, and writing – review & editing. Yuefeng Zhang: investigation, data curation, and writing – original draft. Zimeng Ye: investigation and data curation. Ting Han: investigation and data curation. Honglu Hu: data curation. Ruijie Yang: writing – review & editing. Ting Ying: data curation. Weikang Zheng: data curation. Ruixin Yan: data curation. Yue Zhang: data curation. Zhenbin Wang: methodology and investigation. Zhiyuan Zeng: conceptualization, methodology, writing – review & editing, supervision, project administration, and funding acquisition.

Conflicts of interest

The authors declare that there are no competing interests.

Acknowledgements

Z. Y. Zeng acknowledges the General Research Fund (GRF) support from the Research Grants Council of the Hong Kong Special Administrative Region, China (project no. CityU11308923), the Basic Research Project from Shenzhen Science and Technology Innovation Committee in Shenzhen, China (no. JCYJ20210324134012034), the Applied Research Grant of City University of Hong Kong (project no. 9667247) and the Chow Sang Sang Group Research Fund of City University of Hong Kong (project no. 9229123). Z. Y. Zeng also acknowledges the funding support from the Seed Collaborative Research Fund Scheme of the State Key Laboratory of Marine Pollution, which receives regular research funding from the Innovation and Technology Commission (ITC) of the Hong Kong SAR Government. However, any opinions, findings, conclusions or recommendations expressed in this publication do not reflect the views of the Hong Kong SAR Government or the ITC.

References

  1. A. H. Shah, Z. S. Zhang, Z. H. Huang, S. B. Wang, G. Y. Zhong, C. Z. Wan, A. N. Alexandrova, Y. Huang and X. F. Duan, The role of alkali metal cations and platinum-surface hydroxyl in the alkaline hydrogen evolution reaction, Nat. Catal., 2022, 5, 923–933 CrossRef CAS.
  2. S. Chu and A. Majumdar, Opportunities and challenges for a sustainable energy future, Nature, 2012, 488, 294–303 CrossRef CAS PubMed.
  3. Z. Y. Yu, Y. Duan, X. Y. Feng, X. X. Yu, M. R. Gao and S. H. Yu, Clean and Affordable Hydrogen Fuel from Alkaline Water Splitting: Past, Recent Progress, and Future Prospects, Adv. Mater., 2021, 33, 2007100 CrossRef CAS PubMed.
  4. A. Sartbaeva, V. L. Kuznetsov, S. A. Wells and P. P. Edwards, Hydrogen nexus in a sustainable energy future, Energy Environ. Sci., 2008, 1, 79–85 RSC.
  5. H. P. Xie, Z. Y. Zhao, T. Liu, Y. F. Wu, C. Lan, W. C. A. Jiang, L. Y. Zhu, Y. P. Wang, D. S. Yang and Z. P. Shao, A membrane-based seawater electrolyser for hydrogen generation, Nature, 2022, 612, 673–678 CrossRef CAS PubMed.
  6. G.-H. Gao, R.-Z. Zhao, Y.-J. Wang, X. Ma, Y. Li, J. Zhang and J.-S. Li, Core-shell heterostructure engineering of CoP nanowires coupled NiFe LDH nanosheets for highly efficient water/seawater oxidation, Chin. Chem. Lett. DOI:10.1016/j.cclet.2023.109181.
  7. D. Strmcnik, P. P. Lopes, B. Genorio, V. R. Stamenkovic and N. M. Markovic, Design principles for hydrogen evolution reaction catalyst materials, Nano Energy, 2016, 29, 29–36 CrossRef CAS.
  8. S. Yang, J. Y. Zhu, X. N. Chen, M. J. Huang, S. H. Cai, J. Y. Han and J. S. Li, Self-supported bimetallic phosphides with artificial heterointerfaces for enhanced electrochemical water splitting, Appl. Catal., B, 2022, 304, 120914 CrossRef CAS.
  9. S. S. Liu, L. J. Ma and J. S. Li, Dual-metal-organic-framework derived CoP/MoP hybrid as an efficient electrocatalyst for acidic and alkaline hydrogen evolution reaction, J. Colloid Interface Sci., 2023, 631, 147–153 CrossRef CAS PubMed.
  10. J. Zhu, L. S. Hu, P. X. Zhao, L. Y. S. Lee and K. Y. Wong, Recent Advances in Electrocatalytic Hydrogen Evolution Using Nanoparticles, Chem. Rev., 2020, 120, 851–918 CrossRef CAS PubMed.
  11. D. Li, X. F. Chen, Y. Z. Lv, G. Y. Zhang, Y. Huang, W. Liu, Y. Li, R. S. Chen, C. Nuckolls and H. W. Ni, An effective hybrid electrocatalyst for the alkaline HER: Highly dispersed Pt sites immobilized by a functionalized NiRu-hydroxide, Appl. Catal., B, 2020, 269, 118824 CrossRef CAS.
  12. X. J. Niu, Y. J. Wang, G. H. Gao, T. D. Yang, J. W. Mei, Y. C. Qi, R. Z. Tian and J. S. Li, Interfacial engineering of CoP/CoS2 heterostructure for efficiently electrocatalytic pH-universal hydrogen production, J. Colloid Interface Sci., 2023, 652, 989–996 CrossRef CAS PubMed.
  13. S. Sarkar and S. C. Peter, An overview on Pd-based electrocatalysts for the hydrogen evolution reaction, Inorg. Chem. Front., 2018, 5, 2060–2080 RSC.
  14. X. Y. Chia and M. Pumera, Characteristics and performance of two-dimensional materials for electrocatalysis, Nat. Catal., 2018, 1, 909–921 CrossRef CAS.
  15. Z. P. Wu, X. F. Lu, S. Q. Zang and X. W. Lou, Non-Noble-Metal-Based Electrocatalysts toward the Oxygen Evolution Reaction, Adv. Funct. Mater., 2020, 30, 1910274 CrossRef CAS.
  16. H. Y. Jin, C. X. Guo, X. Liu, J. L. Liu, A. Vasileff, Y. Jiao, Y. Zheng and S. Z. Qiao, Emerging Two-Dimensional Nanomaterials for Electrocatalysis, Chem. Rev., 2018, 118, 6337–6408 CrossRef CAS PubMed.
  17. Q. Fu, J. C. Han, X. J. Wang, P. Xu, T. Yao, J. Zhong, W. W. Zhong, S. W. Liu, T. L. Gao, Z. H. Zhang, L. L. Xu and B. Song, 2D Transition Metal Dichalcogenides: Design, Modulation, and Challenges in Electrocatalysis, Adv. Mater., 2021, 33, 1907818 CrossRef CAS PubMed.
  18. L. Mei, Q. Y. Zhang, M. Du and Z. Y. Zeng, Electrochemical biosensing platforms on the basis of reduced graphene oxide and its composites with Au nanodots, Analyst, 2020, 145, 3749–3756 RSC.
  19. R. Yang, Y. Fan, L. Mei, H. S. Shin, D. Voiry, Q. Lu, J. Li and Z. Y. Zeng, Synthesis of atomically thin sheets by the intercalation-based exfoliation of layered materials, Nat. Synth., 2023, 2, 101–118 CrossRef.
  20. R. J. Yang, Y. Y. Fan, Y. F. Zhang, L. Mei, R. S. Zhu, J. Q. Qin, J. G. Hu, Z. X. Chen, Y. H. Ng, D. Voiry, S. Li, Q. Y. Lu, Q. Wang, J. C. Yu and Z. Y. Zeng, 2D Transition Metal Dichalcogenides for Photocatalysis, Angew. Chem., Int. Ed., 2023, 62, e202218016 CrossRef CAS PubMed.
  21. X. Wang, J. Wu, Y. W. Zhang, Y. Sun, K. K. Ma, Y. Xie, W. H. Zheng, Z. Tian, Z. Kang and Y. Zhang, Vacancy Defects in 2D Transition Metal Dichalcogenide Electrocatalysts: From Aggregated to Atomic Configuration, Adv. Mater., 2023, 35, 2206576 CrossRef CAS PubMed.
  22. J. Deng, H. B. Li, J. P. Xiao, Y. C. Tu, D. H. Deng, H. X. Yang, H. F. Tian, J. Q. Li, P. J. Ren and X. H. Bao, Triggering the electrocatalytic hydrogen evolution activity of the inert two-dimensional MoS2 surface via single-atom metal doping, Energy Environ. Sci., 2015, 8, 1594–1601 RSC.
  23. L. Mei, X. P. Gao, Z. Gao, Q. Y. Zhang, X. G. Yu, A. L. Rogach and Z. Y. Zeng, Size-selective synthesis of platinum nanoparticles on transition-metal dichalcogenides for the hydrogen evolution reaction, Chem. Commun., 2021, 57, 2879–2882 RSC.
  24. X. Huang, Z. Y. Zeng, S. Y. Bao, M. F. Wang, X. Y. Qi, Z. X. Fan and H. Zhang, Solution-phase epitaxial growth of noble metal nanostructures on dispersible single-layer molybdenum disulfide nanosheets, Nat. Commun., 2013, 4, 1444 CrossRef PubMed.
  25. Z. Y. Luo, J. J. Li, Y. L. Li, D. J. Wu, L. Zhang, X. Z. Ren, C. X. He, Q. L. Zhang, M. Gu and X. L. Sun, Band Engineering Induced Conducting 2H-Phase MoS2 by Pd-S-Re Sites Modification for Hydrogen Evolution Reaction, Adv. Energy Mater., 2022, 12, 2103823 CrossRef CAS.
  26. Z. Y. Luo, Y. X. Ouyang, H. Zhang, M. L. Xiao, J. J. Ge, Z. Jiang, J. L. Wang, D. M. Tang, X. Z. Cao, C. P. Liu and W. Xing, Chemically activating MoS2 via spontaneous atomic palladium interfacial doping towards efficient hydrogen evolution, Nat. Commun., 2018, 9, 2120 CrossRef PubMed.
  27. Z. X. Chen, K. Leng, X. X. Zhao, S. Malkhandi, W. Tang, B. B. Tian, L. Dong, L. R. Zheng, M. Lin, B. S. Yeo and K. P. Loh, Interface confined hydrogen evolution reaction in zero valent metal nanoparticles-intercalated molybdenum disulfide, Nat. Commun., 2017, 8, 14548 CrossRef CAS PubMed.
  28. Z. Y. Wang, A. Sim, J. J. Urban and B. X. Mi, Removal and Recovery of Heavy Metal Ions by Two-dimensional MoS2 Nanosheets: Performance and Mechanisms, Environ. Sci. Technol., 2018, 52, 9741–9748 CrossRef CAS PubMed.
  29. F. Q. Liu, S. J. You, Z. Y. Wang and Y. B. Liu, Redox-Active Nanohybrid Filter for Selective Recovery of Gold from Water, ACS ES&T Eng., 2021, 1, 1342–1350 Search PubMed.
  30. L. Mei, Y. Zhang, T. Ying, W. Zheng, H. Hu, R. Yang, R. Yan, Y. Zhang, C. Cheng, B. Liu, S. Li and Z. Y. Zeng, Photochemical reduction of ultrasmall Pt nanoparticles on single-layer transition-metal dichalcogenides for hydrogen evolution reactions, Mater. Today Energy, 2024, 42, 101487 CrossRef CAS.
  31. Z. Y. Zeng, Z. Y. Yin, X. Huang, H. Li, Q. Y. He, G. Lu, F. Boey and H. Zhang, Single-Layer Semiconducting Nanosheets: High-Yield Preparation and Device Fabrication, Angew. Chem., Int. Ed., 2011, 50, 11093–11097 CrossRef CAS PubMed.
  32. R. J. Yang, L. Mei, Q. Y. Zhang, Y. Y. Fan, H. S. Shin, D. Voiry and Z. Y. Zeng, High-yield production of mono- or few-layer transition metal dichalcogenide nanosheets by an electrochemical lithium ion intercalation-based exfoliation method, Nat. Protoc., 2022, 17, 358–377 CrossRef CAS PubMed.
  33. G. Kresse and J. Furthmuller, Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set, Phys. Rev. B: Condens. Matter Mater. Phys., 1996, 54, 11169–11186 CrossRef CAS PubMed.
  34. G. Kresse and J. Hafner, Ab initio molecular dynamics for liquid metals, Phys. Rev. B, 1993, 47, 558 CrossRef CAS PubMed.
  35. J. P. Perdew, K. Burke and M. Ernzerhof, Generalized gradient approximation made simple, Phys. Rev. Lett., 1996, 77, 3865–3868 CrossRef CAS PubMed.
  36. S. Grimme, J. Antony, S. Ehrlich and H. Krieg, A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu, J. Chem. Phys., 2010, 132, 154104 CrossRef PubMed.
  37. S. Maintz, V. L. Deringer, A. L. Tchougréeff and R. Dronskowski, LOBSTER: A Tool to Extract Chemical Bonding from Plane-Wave Based DFT, J. Comput. Chem., 2016, 37, 1030–1035 CrossRef CAS PubMed.
  38. R. F. W. Bader, ATOMS IN MOLECULES, Acc. Chem. Res., 1985, 18, 9–15 CrossRef CAS.
  39. J. K. Nørskov, J. Rossmeisl, A. Logadottir, L. Lindqvist, J. R. Kitchin, T. Bligaard and H. Jónsson, Origin of the overpotential for oxygen reduction at a fuel-cell cathode, J. Phys. Chem. B, 2004, 108, 17886–17892 CrossRef.
  40. W. B. Li, X. F. Qian and J. Li, Phase transitions in 2D materials, Nat. Rev. Mater., 2021, 6, 829–846 CrossRef CAS.
  41. T. Kim, T. H. Lee, S. Y. Park, T. H. Eom, I. Cho, Y. Kim, C. Kim, S. A. Lee, M. J. Choi, J. M. Suh, I. S. Hwang, D. Lee, I. Park and H. W. Jang, Drastic Gas Sensing Selectivity in 2-Dimensional MoS2 Nanoflakes by Noble Metal Decoration, ACS Nano, 2023, 17, 4404–4413 CrossRef CAS PubMed.
  42. Z. Y. Shi, X. Zhang, X. Q. Lin, G. G. Liu, C. Y. Ling, S. B. Xi, B. Chen, Y. Y. Ge, C. L. Tan, Z. C. Lai, Z. Q. Huang, X. Y. Ruan, L. Zhai, L. J. Li, Z. J. Li, X. X. Wang, G. H. Nam, J. W. Liu, Q. Y. He, Z. Q. Guan, J. L. Wang, C. S. Lee, A. R. J. Kucernak and H. Zhang, Phase-dependent growth of Pt on MoS2 for highly efficient H2 evolution, Nature, 2023, 621, 300–305 CrossRef CAS PubMed.
  43. L. Mei, Z. L. Cao, T. Ying, R. J. Yang, H. R. Peng, G. Wang, L. Zheng, Y. Chen, C. Y. Tang, D. Voiry, H. H. Wang, A. B. Farimani and Z. Y. Zeng, Simultaneous Electrochemical Exfoliation and Covalent Functionalization of MoS2 Membrane for Ion Sieving, Adv. Mater., 2022, 34, 2201416 CrossRef CAS PubMed.
  44. J. Li, P. Song, J. P. Zhao, K. Vaklinova, X. X. Zhao, Z. J. Li, Z. Z. Qiu, Z. H. Wang, L. Lin, M. Zhao, T. S. Herng, Y. X. Zuo, W. Jonhson, W. Yu, X. Hai, P. Lyu, H. M. Xu, H. M. Yang, C. Chen, S. J. Pennycook, J. Ding, J. H. Teng, A. H. C. Neto, K. S. Novoselov and J. Lu, Printable two-dimensional superconducting monolayers, Nat. Mater., 2021, 20, 181–187 CrossRef CAS PubMed.
  45. Z. Gao, L. Mei, J. K. Zhou, Y. Fu, L. Zhai, Z. Y. Li, R. J. Yang, D. F. Li, Q. Zhang, J. H. He, J. Li, X. C. Huang, H. Li, Y. M. Liu, K. M. Yao, Y. Y. Gao, L. Zheng, Y. Chen, D. Y. Lei, H. Zhang, Z. Y. Zeng and X. E. Yu, Room-temperature-processed transparent hemispherical optoelectronic array for electronic eyes, Mater. Today, 2023, 69, 31–40 CrossRef CAS.
  46. H. J. Huang, Y. Xue, Y. S. Xie, Y. Yang, L. Yang, H. Y. He, Q. G. Jiang and G. B. Ying, MoS2 quantum dot-decorated MXene nanosheets as efficient hydrogen evolution electrocatalysts, Inorg. Chem. Front., 2022, 9, 1171–1178 RSC.
  47. L. L. Zhang, Q. W. Chang, H. M. Chen and M. H. Shao, Recent advances in palladium-based electrocatalysts for fuel cell reactions and hydrogen evolution reaction, Nano Energy, 2016, 29, 198–219 CrossRef CAS.
  48. X. Zhang, Z. M. Luo, P. Yu, Y. Q. Cai, Y. H. Du, D. X. Wu, S. Gao, C. L. Tan, Z. Li, M. Q. Ren, T. Osipowicz, S. M. Chen, Z. Jiang, J. Li, Y. Huang, J. Yang, Y. Chen, C. Y. Ang, Y. L. Zhao, P. Wang, L. Song, X. J. Wu, Z. Liu, A. Borgna and H. Zhang, Lithiation- induced amorphization of Pd3P2S8 for highly efficient hydrogen evolution, Nat. Catal., 2018, 1, 460–468 CrossRef CAS.

Footnotes

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4qi00622d
These authors contributed equally.

This journal is © the Partner Organisations 2024
Click here to see how this site uses Cookies. View our privacy policy here.