One-step hydrothermal synthesis of Fe single atom doped 1T-MoS2 nanosheets for high-performance seawater hydrogen production
Abstract
In the pursuit of sustainable and clean energy sources, the development of efficient electrocatalysts for hydrogen evolution reaction has gained significant attention. In this work, we synthesized single-atom Fe-doped 1T MoS2 (sFe-1T/MoS2) nanosheets using a one-step hydrothermal method, harnessing the synergistic effects of iron-intercalation to enhance hydrogen production through an abundance of active sites. Notably, 10 at.% sFe-1T/MoS2 exhibited excellent hydrogen evolution reaction performance with a low onset potential of 190 mV and a Tafel slope of 55 mV/dec in acidic solution. High performance was also achieved in alkaline solutions. Additionally, these catalysts demonstrated excellent efficiency in seawater splitting. This work not only offers a cost-effective and scalable method for producing high-quality electrocatalysts but also sets a precedent for the application of this technology across various catalytic systems, marking a significant advancement in clean energy research.
Keywords
INTRODUCTION
The search for clean energy sources remains the most critical challenge of our time, especially in light of serious consequences induced by greenhouse gas emissions[1]. Hydrogen gas is widely recognized as the cleanest alternative to fossil fuels, as its combustion only produces water. Traditional methods to produce hydrogen are not environmentally friendly due to significant CO2 emissions. Catalysis is promising to achieve the production of hydrogen with minimal or no greenhouse emissions utilizing processes such as photocatalysis, electrocatalysis, photothermal catalysis, etc.[2-5]. Among these, electrocatalysis, a process that produces hydrogen from water using catalysts with a small amount of energy input, has shown great potential for hydrogen clean energy. However, achieving efficient and cost-effective hydrogen production remains challenging, as the expensive Pt catalyst is still the best catalyst due to its outstanding water-splitting efficiency during electrocatalysis (Tafel slope: 30 mV/dec). However, its high cost and scarcity limit its wide applications. Efforts have been made to search for other catalysts as the working electrodes for efficient water splitting, such as graphene, MXene, porous carbons, molybdenum disulfide (MoS2), etc.[6,7]. Among them, MoS2 has gained considerable attention due to its high conductivity, low synthesis cost and ultra stability during electrocatalysis. MoS2 features a stable semiconductive phase (2H) and a metastable metallic phase (1T)[8,9]. The 2H phase has limited active sites, making the water splitting less favorable[10]. Specially, the Tafel reaction in water splitting must overcome a high activation energy of 0.98 eV to initiate the Heyrovsky reaction on 2H-MoS2 basal plane, while 1T phase has a much lower energy barrier of 0.16 eV, followed by a smaller Heyrovsky reaction activation energy (0.62 eV) due to the vibration (S-H edge) at sulfur site in the 1T phase, resulting in a higher hydrogen generation rate on the basal edge of 1T MoS2[11-14]. However, 1T phase MoS2 is a metastable phase and achieving long term stable hydrogen production is still challenging. Moreover, its water-splitting performance still has a distance to Pt metals. Therefore, further research is necessary to enhance the performance of 1T-phase MoS2 by advanced material strategies, such as doping, intercalation, surface modification, etc.[3].
Strain engineering has emerged as an effective strategy for stabilizing metastable structures. Our previous work successfully stabilized 1T MoS2 with hydrazine intercalation. The as-synthesized 1T dominant MoS2 nanosheets exhibit significantly improved hydrogen evolution reaction (HER) performance (64 mV/dec) with increased active sites, which makes 1T-MoS2 a bright future for highly efficient water production[15]. Doping, another widely used efficient strategy, introduces multiple lattice strains and defect centers through surface modification. Single-atom doping in electrocatalysis has shown a promising future in terms of catalytic activity, selectivity and stability due to the increased active sites on the catalyst surface[16]. By selecting the appropriate dopants, the electronic properties of the host material can be tailored to optimize catalytic performance. Transition metal featuring a high d-orbital status shows a similar effect to Pt on electrocatalysis[17], especially Ni, Cu, V and W, as reported in previous millstones [Supplementary Table 1], showing advanced water splitting efficiency. For example, by doping 5% V, the doped 1T-MoS2 exhibited a low onset potential of 102 mV and a Tafel slope of 54 mV/dec[7]. The results indicated that doping with transition metals in MoS2 could effectively modulate the energy level, which decreases the energy barrier for electron transfer and consequently reduces the overpotential. The performance and stability of 1T-MoS2 can also be improved by intercalating single atoms into its layers. For example, a high concentration of single atoms of copper (98 at.%) was successfully intercalated into the layers of 1T MoS2 by simple stirring. The synthesized Cu-1T MoS2 optimal sample showed a low onset potential of 104 mV and a small Tafel slope of 52 mV/dec with a high stability and the 1T phase can be maintained up to 800 oC. The enhanced catalytic performance was attributed to improved electrical conductivity, enhanced active sites, and modified electronic structure with weakened H-S bond[18]. Another example includes the work done by Zhao et al., who incorporated FeS2 clusters with MoS2, demonstrating a high HER performance with a Tafel slope of 82 mVdec-1 and onset potential of 136 mV at a current density of 10 mAcm-2[19].
The advancement of electrocatalysts for HER in seawater splitting represents a significant leap toward sustainable hydrogen production. The high efficiency achieved through active ion concentration manipulation underscores the potential for scaling this methodology for industrial applications. Fe is one of the promising transition metals for efficient water splitting due to the S2p-Fe3d interaction[20]. In this work, we develop a single-atom system of Fe-doped 1T MoS2 (sFe-1T/MoS2) nanosheets, a material that has recently gained attention for its promising hydrogen electrocatalysis due to the S2p-Fe3d interaction. Notably, the synthesized sFe-1T/MoS2 nanosheets exhibit a unique structural evolution from single-atom intercalation to site-replacement doping, a phenomenon not extensively documented in previous studies. This structural transformation by doping contributes to a remarkable electrochemical performance, particularly for the 10 at.% sFe-1T/MoS2, which achieves an onset potential of 190 mV and a Tafel slope of 55 mV/dec in acidic solutions. In alkaline solutions, the material displays an extraordinarily low onset potential of 80 mV alongside a slightly higher Tafel slope compared to that in acidic solutions. This dual-performance characteristic in varying pH environments highlights the material's versatility and robustness. The significant enhancement in catalytic activity is attributed to the morphological effects that induced more active sites due to the change in the coordination environment, as confirmed through advanced Near Edge X-ray Absorption Fine Structure (NEXAFS) spectroscopy. Moreover, the synthesized sFe-1T/MoS2 demonstrates exceptional performance in ocean water splitting, an area of growing interest due to the increasing focus on sustainable and scalable hydrogen production methods. This research provides a novel, facile, and cost-effective strategy for developing highly efficient electrocatalysts, which could revolutionize the field of sustainable energy. Our work not only underscores the potential of sFe-1T/MoS2 in various electrochemical applications but also paves the way for further exploration of single-atom doped catalysts. The insights gained from this study could inspire the development of next-generation catalytic systems with enhanced performance and durability.
EXPERIMENTAL SECTION
Materials
Fe-doped MoS2 was synthesized by a hydrothermal reaction. All those chemicals were purchased from Sigma-Aldrich without any further purification. Ammonium heptamolybdate [(NH4)6Mo7O24 · 4H2O] was used as the Mo source (sMo) since it matches the microstructure of 1T-MoS2 because of its fine Mo-centre octahedral structure. Fe(NO3)3 ∙ 9H2O was chosen as the Fe-dopant and the doping concentration varied in a ratio of 2 at.%, 5 at.%, 10 at.% and 20 at.%, respectively. First, 10 mmol sulfur powders and 5 mmol sMo were mixed and fully dissolved in 10 ml hydrazine monohydrate (N2H4 · H2O) via 20 min ultrasonication. The Fe-dopant was dissolved in 50 ml of deionized (DI) water in another beaker. Then, those two precursors were mixed in 80 mL stainless steel autoclaves for 48 h at 180 oC. The impurities in the product were fully washed using 0.5 M H2SO4 and DI water, accompanied by 5000 rpm centrifugation for 15 min. Then, the wet sample was transferred to a vacuum oven for drying for 48 h at room temperature. The pristine 1T-MoS2 is marked as S0 as the reference compared to the Fe-doped MoS2 labeled as S2 (2 at.%), S5 (5 at.%), S10 (10 at.%) and S20 (20 at.%), respectively. The acidic corrosion to iron-doped MoS2 was also checked by acidic treatment (2M HCL wash). The prepared samples (800 mg) were put in 50 ml 2M HCL solution with 5 mins stirring. The washed samples were dried in the same way as the sample preparation.
Characterization
The surface morphology was investigated by a JEOL scanning electron microscope (JEOL-SEM) at room temperature. The crystal structure of sFe-1T/MoS2 was investigated by a JEOL F200 Transmission Electron Microscope (F200-TEM) at 200 kV voltage. X-ray diffraction analysis (XRD) was carried out by a Panalytical Empyrean instrument with λ = 1.54 Å (CuKα) that generates at 40 kV voltage and 40 mA current. The sample was loaded on a circular holder and the angle of measurements is within a 2θ scan range of 5° to 60°. The scan step size is 0.006° with a time-per-step of 97.92s. Raman measurement was carried out on a Witec alpha-300 spectrometer. A 532 nm laser source was used as the excitation source for 1 h measurement under 0.1% laser power. The textural Brunauer-Emmett-Teller (BET) surface area was determined by the N2 adsorption-desorption analysis conducted on the Micromeritics ASAP 2420 instrument. The samples were degassed under a constant vacuum at 200 °C before analysis. The degassed samples were analyzed at the liquid nitrogen temperature (-196 °C). The BET method was used to calculate the total specific surface area. Total pore volume was obtained at a relative pressure (P/P0) of 0.99. The micropore characteristics, including the micropore surface area and the micropore volume, were obtained using the t-plot method. The coordination information was also obtained by the low-energy (soft) and high-energy (hard) near-edge X-ray absorption fine structure (NEXAFS) analysis in a total fluorescence yield (TFY) mode at the Synchrotron of Singapore University and the Australia Synchrotron, respectively. X-ray Photoelectron Spectroscopy (XPS) and Ultraviolet Photoelectron Spectroscopy (UPS) were measured by the SPECS Phobis 100 in Chengdu, China. The UPS is performed in He (I) spectrum. All measurements above were taken at room temperature.
Measurements
The electrochemical measurement was carried out using a three-electrode system on CHI 760E electrochemical workstation. The working electrode is the as-deposit samples on glassy-carbon rod electrode (sample loading amount ~ 40 µg·cm-2), Ag/AgCl (3.5 M KCl) and Pt metal was applied as the reference electrode and counter electrode, respectively. 30 mL of 0.5M H2SO4 and 1M KOH were used separately as acidic and alkaline electrolytes. Linear sweep voltammetry (LSV) and chronoamperometry were applied with a scan rate of 5.0 mV s-1. The measurements were performed from -0.1 V to -0.8 V (vs. Ag/AgCl) in 0.5 M H2SO4 and from -0.9 V to -1.6 V (vs. Ag/AgCl) in 1.0 M KOH. The LSV performance was also checked in the seawater environment which was picked from the eastern coastline of Newcastle Beach, Australia. All the recorded potentials were then calibrated vs. reversible hydrogen electrode (RHE) based on the Nernst equation:
The electrochemical impedance spectroscopy (EIS) measurement was carried out at an overpotential of
RESULTS AND DISCUSSION
The XRD patterns of sFe-1T/MoS2 are presented in Figure 1A. The broad peak at 9.5o indicates the (002) orientation of 1T/MoS2. Compared to the 2H domain peak at 17.5o, the left-shift peak position indicates the structure change from 2H hexagonal to 1H monolayers, as described in previous works[4,21,22]. The peak position moves slightly to a lower angle when the dopant concentration increases, which indicates the enlarged interlayer space caused by Fe doping. This trend is consistent with our previous work on the Cu single-atom system and suggests the possible formation of single-atom Fe coordination. The impurity peaks in sample S20 indicate the formation of FeSx clusters (surface aggregation), which is beyond the ideal concentration of iron doping. Raman measurement was also employed to investigate the phase variation of sFe-1T/MoS2. As displayed in Figure 1B, 1T- dominant bond vibration is confirmed by intense peaks at the Raman shift of 147 and 324 cm-1 for 1T-phase J1 (out of plane) and J3 (in-plane) Mo-S phonon modes, respectively[23]. As the dopant concentration increases, a leftward shift of the J1 peak (J1 → J1*) with a broadened peak area (S0 → S10) indicates changes in anisotropic symmetry due to Fe doping, while a J2 mode is also presented in sample S20. Generally, the J2 vibrational mode is related to in-plane Mo-S displacements (defects). This indicates weak zigzag (S-Mo-S) displacements at low doping concentrations supported by single-atom intercalation[24]. While the strong site replacement effect is observed at high doping concentrations (S20) instead of layer intercalation. Additionally, peak splitting is also observed for the in-plane J3 vibrational mode (J3 → J3* and J3 → J3’, respectively) and the J3 peak suddenly disappears in sample S20. The pronounced peak splitting highlights the twofold symmetric variation between single-atom doping and site replacement doping and also indicates doping-induced phase variation from 1T to 2H[25]. Furthermore, the peak at 407 cm-1 corresponds to the out-of-plane S-Mo-S opposite moving (A1g) of MoS2. The peak intensity of A1g mode decreases without a noticeable peak position shift, suggesting a weakening of out-of-plane S-Mo-S vibrations due to a formation of S-Fe-S bulks (phase aggregation), agreeing with the SEM results described above. The absence of the E1g mode is primarily attributed to low laser density as a laser-induced 1T to 2H phase transition can be triggered under high laser density conditions[26]. The deep lattice pattern is further demonstrated by transmission electron microscopy (TEM) from sample S0 to sample S10. The high-resolution TEM image of sample S10 is displayed in Figure 1C and that of others is demonstrated in Supplementary Figure 1. The apparent wrinkles indicate the fine sheet structure of sFe-1T/MoS2 at low doping concentrations. Some Fe atoms are found between interlayers of MoS2 without prominent Fe clusters, suggesting the formation of single atoms. In addition, there is a plane expansion (002) in doped MoS2 1T-dominance, from 0.83 nm to 0.98 nm, when the doping concentration increases from 0 to 10 at.%. Those observations prove the Fe single-atoms intercalation between MoS2 layers, consistent with our previous works on Cu intercalated 1T/MoS2[18].
Figure 1. The characterization of sFe-1T/MoS2. (A) XRD patterns; (B) Raman spectra; (C) High-resolution TEM image of sample S10; (D) XANES curves of Fe L3-edge; (E) NEXAFS curves of Fe K-edge with Fe foil reference; (F) Fourier-transform (FT) k2-weighted NEXAFS spectra of sFe-1T/MoS2 with Fe foil reference; Wavelet transform (WT) of k2-weighted NEXAFS spectra for (G) S2, (H) S5, (I) S10 and (J) S20 based on the mode of Morlet wavelet function (κ = 3; σ = 0.5) for the first coordination shells, where 1 ≤ (R+α) ≤ 3 Å and 2 ≤ k ≤ 10 Å-1. XRD: X-ray diffraction analysis; TEM: Transmission electron microscope; NEXAFS: Near edge X-ray absorption fine structure; XANES: X-ray absorption near edge structure.
To understand the electron environment of iron intercalation, the L3-edge orbital status of iron is explained by low-energy (soft) NEXAFS measurement to determine the dynamic process of p-d dipole excitation, as demonstrated in Figure 1D. The 2p → 3d electron transfer occurs between 708 eV and 710 eV without obvious t2g-eg orbital splitting in all samples. Additionally, only an L3 p-d electron transfer peak is observed at 709 eV, indicating the absence of strong crystal field splitting effect. This behavior is mainly attributed to the low spin state (unoccupied eg dominant) of Fe2+[27]. A slightly right shift of peak position (S2 → S10) indicates a lower coordination environment and higher oxidization state since a lower valence band position can be predicted[28]. However, there is no shift for sample S20, which is associated with the structural aggregation mentioned above. The peak intensity increases and the peak width also broadens symmetrically aligned with the rise in Fe concentrations (S2 → S10). Generally, the higher peak intensity indicates a higher unoccupied density state (more empty orbitals) and lower covalency caused by higher dopant concentration[29]. In contrast, the enlarged peak width indicates the expanded unoccupied orbitals and enhanced orbital contributions caused by multi-coordination[30]. It is observed that there is no obvious post-edge shift from samples S10 to S20, while a pre-edge shift is noticeable. This indicates the new subband formed aligning with the transition from a low spin multiplet (unoccupied eg dominant) to high spin multiplet (t2g/eg equal contribution) transition when the Fe concentration keeps increasing (new phase generated)[31]. Fe K-edge (1s → 4p dipole electron transition) is further checked by high energy (hard) NEXAFS and the normalized adsorption edge is displayed in Figure 1E. Generally, the shift of edge position is associated with the change of oxidization state, such as the transition from ferrous to ferric. All samples show a higher edge position than the reference curve of Fe foil, which confirms the similar oxidization state (Fe2+) of iron in sFe-1T/MoS2. A pre-edge peak at 7,113 eV for sample S20 is observed compared to a flat appearance in samples with lower doping concentrations. The pre-edge feature originates from a quadrupole (1s to 3d) transition, indicating a 3d→4p geometric coordination hybridization from octahedral to tetrahedral (Td) symmetry, accordingly[32]. It also has a rising edge level (7,115 ~ 7,120 eV) from S2 to S20, while this edge lifting between 7,115 eV and 7,120 eV, in our case, is mainly affected by the pre-edge feature aligning with the increased dopant concentration since there is no noticeable edge position shift to Fe-doped samples. The feature peak at 7,120 eV in sample S20 indicates the bulk formation of FeS2. This is confirmed by SEM images [Supplementary Figure 2]. The nanoflower appearance (S0 → S10) indicates fine nanosheet structure of MoS2, shrinking in dry conditions, while a bulk-like morphology is observed in sample S20. This proves a sheet to bulk transition with increased doping concentration, which is widely observed in other doping systems. The intensity of the adsorption peak at 7,127 eV decreases, accompanied by an elevated peak intensity at 7,135 eV when dopant concentration grows to 5 at.%, which corresponds to the standard FeS to FeS2 transition[33]. A slightly decreased peak intensity from samples S5 to S20 is mainly caused by orbital distortion, corresponding to the pre-edge changes. The distance of the first atom shell is obtained by Fourier k2-weighted transform (k2-FT), as shown in Figure 1F. Compared to the Fe-Fe scattering path in Fe foil (2.2 Å), the absence of Fe–Fe interaction indicates the single-atom configuration for sFe-1T/MoS2. The peak at approximately 1.7 Å is mainly attributed to the first shell Fe-S scattering path coordinated at the S site between the basal planes[34]. In addition, there is a slight peak shifting to a shorter scattering path to 1.5 Å with the increased Fe concentration where a decreased peak intensity of the Fe-S distribution is aligned, which is associated with a reduced Fe-S coordination number and the oscillation from octahedral to tetrahedral coordination environment[35]. The reliability of Fourier transform (FT) is further confirmed by the forward-backward FT (k → R → q). The good matching between k-space and
The catalysis performance for HER was investigated by three-electrode LSV measurement. As displayed in Figure 2A, the HER performance of the sFe-1T/MoS2 (S2, S5 and S10) is gradually improved by the increased dopant concentration compared to pristine 1T/MoS2 nanosheets in an acidic solution. Iron-doped samples exhibit better HER performances with decreased overpotentials. In particular, sample S10 shows the best performance with an onset potential of 190 mV vs. RHE and an overpotential of 215 mV vs. RHE
Figure 2. Electrochemical measurements for all Fe-1T/MoS2 samples. (A) LSV measurements in 0.5 M H2SO4 solution; (B) Tafel slopes derived from LSV curves; (C) Fitting plots of Cdl from CV measurements for the EASA calculation; (D) The EIS in 0.5 M H2SO4 solution; (E) The stability measurements for the acidic and alkaline solution for sample S10; (F) 24 h stability measurement at a constant RHE potential of - 0.2 V (equal to 10 mA/cm2 current density in acidic (0.5 M H2SO4) and alkaline (1M KOH) solutions, respectively. LSV: Linear sweep voltammetry; RHE: Reversible hydrogen electrod; CV: Cyclic voltammetry; EASA: Electrochemically active surface area; EIS: Electrochemical impedance spectra.
To investigate the mechanism of the enhanced HER, the valence band structure was examined by UPS (HeI) measurement and the overall spectra are displayed in Supplementary Figure 10. There is no obvious drift in ionization energy for all 1T-MoS2 samples, indicating that the enhancement of HER is related to more active sites induced by coordination environment change instead of doping-related work function variation[40]. The zoom-in spectra of UPS are demonstrated in Figure 3A to check the variation in 2p orbitals. In the Fe single atom system, the peak of 2p orbitals shifts slightly to lower binding energy from
Figure 3. (A)-(B) UPS (HeI) spectra of s-Fe-1T/MoS2; (C) Coordination number changes with increased Fe concentration; (D) Schematic diagram of catalysis process of sFe-1T/MoS2. UPS: Ultraviolet photoelectron spectroscopy.
To further validate our findings, we applied two additional RHE potentials of -0.3 V and -0.5 V during the stability test. The X-ray absorption spectroscopy (XAS) results for sample S10 after catalysis are presented in Supplementary Figure 12 in the supporting information. As shown in Supplementary Figure 12A, there are no significant changes in the pre-edge peak before and after catalysis at an RHE potential of -0.3 V, indicating minimal variation in the coordination environment after catalysis. However, an increase in the quadrupole pre-edge peak intensity was observed with prolonged catalysis at -0.5V, suggesting a shift towards a higher electronegativity and a more asymmetric coordination environment. The edge position also shifts to a higher energy range, from 7,120 eV to 7,125 eV, reflecting an increase in the Fe oxidation state and the formation of a higher electronegativity center, consistent with the pre-edge changes. Additionally, a new merged white peak at 7,130 eV emerges post-catalysis compared to the split peak observed before catalysis, indicating a newly formed dominant p orbital contribution. This p orbital contribution diminishes when the catalysis was performed at high potential, which suggests the S-site peeling off at high-voltage catalysis. The R-space curves, displayed in Supplementary Figure 12B, further support these observations. A reduction in atomic distance from approximately 1.8 Å to 1.4 Å is observed, along with a decrease in peak intensity. It suggests a reduced coordination number of the Fe site with an indicative transition from FeSx to FeOx.
More active sites induced by iron single atoms also promote significant enhancement in real industrial applications, such as ocean water splitting. In normal ocean water, lower OH-/H+ concentration (neutral pH) leads to a low catalytical efficiency. Thus, to evaluate the efficiency of ocean water splitting, 10% v/v and 30% v/v 0.5M H2SO4 and 1M KOH solutions were added to the ocean water, respectively. The performance in pure acidic and alkaline solutions is remarked as S-H100 and S-OH100, respectively, as the standard reference. The LSV curves of ocean water splitting are demonstrated in Figure 4A and the overpotential at 10 mAcm-2 has been summarized in Supplementary Table 4 supported by previous milestones in MoS2-based seawater splitting. Tafel slope keeps increasing when pH is close to a neutral level, which is mainly caused by reduced active catalytic ions in the environment. This also indicates the reliability of electro-catalytical performance for sFe-1T/MoS2. The long-term stability was also checked, as displayed in Figure 4B. Catalysts in acidic solution show a similar trend with different H+ concentrations, indicating stable performance of acidic ocean splitting. A serious decay is observed in the initial stage (within 3 h) for alkaline-modified ocean water compared to pure 1M KOH solution. This is mainly caused by the newly formed Ca/Mg-based carbonates in the ocean water, which block the active surface of MoS2 and significantly reduce the performance of water splitting. This also explains why this phenomenon is not obvious in acidic solution since it is corroded immediately in H+ environment. The overpotential at 10 mA/cm2 reaches 242 mV for the S-OH30 sample, which is relatively lower compared to previous MoS2 catalysis. In addition, chloride ions in seawater will also significantly affect the HER by competing with hydrogen ions for adsorption sites on the electrode surface, thereby reducing the efficiency of hydrogen production. This explains the significant decrease in current density observed in seawater compared to pure acidic and alkaline solutions. In addition, the chloride ion in seawater also significantly influences HER performance by competing with hydrogen ions on adsorption sites of electrodes, reducing the efficiency of hydrogen production consequently. This competition results in a notable decrease in current density observed in seawater environment compared to pure acidic and alkaline solutions. Furthermore, Cl- can induce electrode corrosion. Although the corrosive effect in alkaline solutions is less effective than in neutral or acidic environments, chloride-induced sample peeling off is the primary concern since Cl ions can penetrate protective oxide layers on the electrode surface, leading to localized corrosion and subsequent the degradation of the electrode's structural integrity.
Figure 4. (A) Seawater HER performance with different volume ratios (v/v) of 0.5M H2SO4 and 1M KOH. The concentration is 10% v/v (S-H10/S-OH10), 30% v/v (S-H30/S-OH30) and 100% v/v (pure solution, S-H100/S-OH100), respectively; (B) The long-term stability in such a solution is tested at a constant RHE potential of -0.2 V within 24 h. HER: Hydrogen evolution reaction; RHE: Reversible hydrogen electrode.
CONCLUSION
In this work, single-atom iron-modified 1T/MoS2 was synthesized using a scalable hydrothermal method. The synthesized samples showed structure transformation from single-atom intercalation to site replacement doping with increasing doping concentrations. Our results demonstrate that doping concentration can significantly influence the HER performance of the doped 1T-phase MoS2. The 10% at. sFe/1T/MoS2 exhibited superior electrochemical activity toward HER with an onset potential of 190 mV and a Tafel slope of 55 mV/dec in acidic solution. Additionally, it shows an exceptionally low onset potential of 80 mV, albeit with a slightly higher Tafel slope in alkaline solution, indicating distinct HER mechanisms. The substantial enhancement is attributed to the morphological effect-induced rich active sites due to the change of coordination environment. Additionally, the synthesized product also exhibited high efficiency in ocean water splitting with an overpotential of 242 mV at 10mA/cm2 in a 30%v/v mixed KOH/seawater environment. This high efficiency of seawater splitting is primarily attributed to the active ions. Our present work provides a facile and cost-effective strategy for designing efficient electrocatalysts for HER, promising broad applicability of this methodology to other catalytic systems. However, overcoming electrode corrosion and achieving high-power output ocean catalysis remain significant challenges requiring further investigation.
DECLARATIONS
Authors’ contributions
Mainly supervised this project: Yi, J.
Carried out this project: Geng, X.
Contributed to sample characterization: Cao, Y.
Performed TEM measurements: Li, Z.
Assisted with catalysis measurement: Li, M.; Chu, D.
Contributed to the materials characterization and revised the manuscript: Cao, C.; Yu, C.; Chu, X.; Hu, L.; Sun, Y.; Wang, D.; Qiao, L.
Conducted the NEXAS measurements: Yu, X.; Breese, M. B. H.
Conducted the EXAFS measurements: Lee, J. M.
Availability of data and materials
The data that support the findings of this study are available from the corresponding author upon reasonable request.
Financial support and sponsorship
This work was supported by the Australian Research Council (DP220103045). The authors would like to acknowledge the Singapore Synchrotron Light Source for providing the facility necessary for conducting the research. The Laboratory is a National Research Infrastructure under the National Research Foundation, Singapore. Any opinions, findings, conclusions or recommendations expressed in this material are those of the author(s) and do not reflect the views of the National Research Foundation, Singapore.
Conflicts of interest
All authors declared that there are no conflicts of interest.
Ethical approval and consent to participate
Not applicable.
Consent for publication
Not applicable.
Copyright
© The Author(s) 2025.
Supplementary Materials
REFERENCES
1. Liu, Z.; Deng, Z.; Davis, S. J.; Giron, C.; Ciais, P. Monitoring global carbon emissions in 2021. Nat. Rev. Earth. Environ. 2022, 3, 217-9.
2. Fan, P.; He, Y.; Pan, J.; et al. Recent advances in photothermal effects for hydrogen evolution. Chinese. Chem. Lett. 2024, 35, 108513.
3. Chu, X.; Sathish, C. I.; Yang, J. H.; et al. Strategies for Improving the photocatalytic hydrogen evolution reaction of carbon nitride-based catalysts. Small 2023, 19, e2302875.
4. Yadav, A. A.; Hunge, Y. M.; Dhodamani, A. G.; Kang, S. W. Hydrothermally synthesized Ag@MoS2 composite for enhanced photocatalytic hydrogen production. Catalysts 2023, 13, 716.
5. Yadav, A.; Hunge, Y.; Kang, S. Porous nanoplate-like tungsten trioxide/reduced graphene oxide catalyst for sonocatalytic degradation and photocatalytic hydrogen production. Surf. Inter. 2021, 24, 101075.
6. Cui, C.; Xie, J.; Lin, D.; et al. Recent progress in carbon-based materials as catalysts for electrochemical and photocatalytic water splitting. In: Paul R, Etacheri V, Wang Y, Lin CT , Editors. Carbon based nanomaterials for advanced thermal and electrochemical energy storage and conversion. Elsevier; 2019.p.173-200.
7. Li, M.; Cai, B.; Tian, R.; et al. Vanadium doped 1T MoS2 nanosheets for highly efficient electrocatalytic hydrogen evolution in both acidic and alkaline solutions. Chem. Eng. J. 2021, 409, 128158.
8. Wang, D.; Zhang, X.; Bao, S.; Zhang, Z.; Fei, H.; Wu, Z. Phase engineering of a multiphasic 1T/2H MoS2 catalyst for highly efficient hydrogen evolution. J. Mater. Chem. A. 2017, 5, 2681-8.
9. Li, M.; Kuo, Y.; Chu, X.; Chu, D.; Yi, J. MoS2 nanoflower incorporated with Au/Pt nanoparticles for highly efficient hydrogen evolution reaction. Emergent. Mater. 2021, 4, 579-87.
10. Jaramillo, T. F.; Jørgensen, K. P.; Bonde, J.; Nielsen, J. H.; Horch, S.; Chorkendorff, I. Identification of active edge sites for electrochemical H2 evolution from MoS2 nanocatalysts. Science 2007, 317, 100-2.
11. Lin, Y. C.; Dumcenco, D. O.; Huang, Y. S.; Suenaga, K. Atomic mechanism of the semiconducting-to-metallic phase transition in single-layered MoS2. Nat. Nanotechnol. 2014, 9, 391-6.
12. Tang, Q.; Jiang, D. Mechanism of hydrogen evolution reaction on 1T-MoS2 from first principles. ACS. Catal. 2016, 6, 4953-61.
13. Li, W.; Liu, G.; Li, J.; et al. Hydrogen evolution reaction mechanism on 2H-MoS2 electrocatalyst. Appl. Surf. Sci. 2019, 498, 143869.
14. Tang, J.; Huang, J.; Ding, D.; Zhang, S.; Deng, X. Research progress of 1T-MoS2 in electrocatalytic hydrogen evolution. Int. J. Hydrogen. Energ. 2022, 47, 39771-95.
15. Li, M.; Zhou, Z.; Hu, L.; et al. Hydrazine hydrate intercalated 1T-dominant MoS2 with superior ambient stability for highly efficient electrocatalytic applications. ACS. Appl. Mater. Inter. 2022, 14, 16338-47.
16. Lei, Z.; Sathish, C.; Liu, Y.; et al. Single metal atoms catalysts-promising candidates for next generation energy storage and conversion devices. EcoMat 2022, 4, e12186.
17. Levy, R. B.; Boudart, M. Platinum-like behavior of tungsten carbide in surface catalysis. Science 1973, 181, 547-9.
18. Li, M.; Selvarajan, P.; Wang, S.; et al. Thermostable 1T-MoS2 nanosheets achieved by spontaneous intercalation of Cu single atoms at room temperature and their enhanced HER performance. Small. Struct. 2023, 4, 2300010.
19. Zhao, X.; Ma, X.; Lu, Q.; et al. FeS2-doped MoS2 nanoflower with the dominant 1T-MoS2 phase as an excellent electrocatalyst for high-performance hydrogen evolution. Electrochim. Acta. 2017, 249, 72-8.
20. Huang, W.; Zhou, D.; Qi, G.; Liu, X. Fe-doped MoS2nanosheets array for high-current-density seawater electrolysis. Nanotechnology 2021, 32, 415403.
21. Acerce, M.; Voiry, D.; Chhowalla, M. Metallic 1T phase MoS2 nanosheets as supercapacitor electrode materials. Nat. Nanotechnol. 2015, 10, 313-8.
22. Liu, L.; Wu, J.; Wu, L.; et al. Phase-selective synthesis of 1T' MoS2 monolayers and heterophase bilayers. Nat. Mater. 2018, 17, 1108-14.
23. Geng, X.; Sun, W.; Wu, W.; et al. Pure and stable metallic phase molybdenum disulfide nanosheets for hydrogen evolution reaction. Nat. Commun. 2016, 7, 10672.
24. Nayak, A. P.; Pandey, T.; Voiry, D.; et al. Pressure-dependent optical and vibrational properties of monolayer molybdenum disulfide. Nano. Lett. 2015, 15, 346-53.
25. Park, M. J.; Yi, S. G.; Kim, J. H.; Yoo, K. H. Metal-insulator crossover in multilayered MoS2. Nanoscale 2015, 7, 15127-33.
26. Nam, G. H.; He, Q.; Wang, X.; et al. In-plane anisotropic properties of 1T'-MoS2 layers. Adv. Mater. 2019, 31, e1807764.
28. Mohamed, A. Y.; Park, W. G.; Cho, D. Y. Chemical structure and magnetism of FeOx/Fe2O3 interface studied by X-ray absorption spectroscopy. Magnetochemistry 2020, 6, 33.
29. Hocking, R. K.; DeBeer, G. S.; Raymond, K. N.; Hodgson, K. O.; Hedman, B.; Solomon, E. I. Fe L-edge X-ray absorption spectroscopy determination of differential orbital covalency of siderophore model compounds: electronic structure contributions to high stability constants. J. Am. Chem. Soc. 2010, 132, 4006-15.
30. Hocking, R. K.; Wasinger, E. C.; Yan, Y. L.; et al. Fe L-edge X-ray absorption spectroscopy of low-spin heme relative to non-heme Fe complexes: delocalization of Fe d-electrons into the porphyrin ligand. J. Am. Chem. Soc. 2007, 129, 113-25.
31. Baker, M. L.; Mara, M. W.; Yan, J. J.; Hodgson, K. O.; Hedman, B.; Solomon, E. I. K- and L-edge X-ray absorption spectroscopy (XAS) and resonant inelastic X-ray scattering (RIXS) determination of Differential orbital covalency (DOC) of transition metal sites. Coord. Chem. Rev. 2017, 345, 182-208.
32. Westre, T. E.; Kennepohl, P.; Dewitt, J. G.; Hedman, B.; Hodgson, K. O.; Solomon, E. I. A multiplet analysis of Fe K-edge 1s → 3d pre-edge features of iron complexes. J. Am. Chem. Soc. 1997, 119, 6297-314.
33. Kitajou, A.; Yamaguchi, J.; Hara, S.; Okada, S. Discharge/charge reaction mechanism of a pyrite-type FeS2 cathode for sodium secondary batteries. J. Power. Sources. 2014, 247, 391-5.
34. Mo, J.; Wu, S.; Lau, T.; et al. Transition metal atom-doped monolayer MoS2 in a proton-exchange membrane electrolyzer. Mater. Today. Adv. 2020, 6, 100020.
35. Pattengale, B.; Huang, Y.; Yan, X.; et al. Dynamic evolution and reversibility of single-atom Ni(II) active site in 1T-MoS2 electrocatalysts for hydrogen evolution. Nat. Commun. 2020, 11, 4114.
36. Zheng, J.; Wu, S.; Lu, L.; et al. Structural insight into [Fe-S2-Mo] motif in electrochemical reduction of N2 over Fe1-supported molecular MoS2. Chem. Sci. 2020, 12, 688-95.
37. Ye, J.; Zang, Y.; Wang, Q.; et al. Nitrogen doped FeS2 nanoparticles for efficient and stable hydrogen evolution reaction. J. Energy. Chem. 2021, 56, 283-9.
38. Solati, N.; Karakaya, C.; Kaya, S. Advancing the understanding of the structure-activity-durability relation of 2D MoS2 for the hydrogen evolution reaction. ACS. Catal. 2023, 13, 342-54.
39. Singh, A. K.; Prasad, J.; Azad, U. P.; et al. Vanadium doped few-layer ultrathin MoS2 nanosheets on reduced graphene oxide for high-performance hydrogen evolution reaction. RSC. Adv. 2019, 9, 22232-9.
40. Shi, Y.; Zhou, Y.; Yang, D. R.; et al. Energy level engineering of MoS2 by transition-metal doping for accelerating hydrogen evolution reaction. J. Am. Chem. Soc. 2017, 139, 15479-85.
41. Hu, J.; Wu, J.; Zhang, S.; et al. One-pot Fabrication of 2D/2D CdIn2S4/In2S3 heterojunction for boosting photocatalytic Cr(VI) reduction. Catalysts 2023, 13, 826.
42. Geng, X.; Singh, G.; Sathish, C.; et al. Biomass derived nanoarchitectonics of porous carbon with tunable oxygen functionalities and hierarchical structures and their superior performance in CO2 adsorption and energy storage. Carbon 2023, 214, 118347.
43. Ying, T.; Yu, T.; Qi, Y.; Chen, X.; Hosono, H. High entropy van der waals materials. Adv. Sci. 2022, 9, e2203219.
44. Kozdra, S.; Wójcik, A.; Das, T.; Michałowski, P. P. From DFT investigations of oxygen-implanted molybdenum disulfide to temperature-induced stabilization of MoS2/MoO3 heterostructure. Appl. Sur. Sci. 2023, 631, 157547.
45. Zander, J.; Timm, J.; Weiss, M.; Marschall, R. Light-induced ammonia generation over defective carbon nitride modified with pyrite. Adv. Energy. Mater. 2022, 12, 2202403.
46. Wolski, S. C.; Kuper, J.; Hänzelmann, P.; et al. Crystal structure of the FeS cluster-containing nucleotide excision repair helicase XPD. PLoS. Biol. 2008, 6, e149.
Cite This Article
How to Cite
Geng, X.; Cao, Y.; Li, Z.; Li, M.; Cao, C.; Yu, C.; Chu, X.; Hu, L.; Sun, Y.; Qiao, L.; Yu, X.; Breese, M. B. H.; Lee, J. M.; Wang, D.; Chu, D.; Yi, J. One-step hydrothermal synthesis of Fe single atom doped 1T-MoS2 nanosheets for high-performance seawater hydrogen production. Energy Mater. 2025, 5, 500009. http://dx.doi.org/10.20517/energymater.2024.46
Download Citation
Export Citation File:
Type of Import
Tips on Downloading Citation
Citation Manager File Format
Type of Import
Direct Import: When the Direct Import option is selected (the default state), a dialogue box will give you the option to Save or Open the downloaded citation data. Choosing Open will either launch your citation manager or give you a choice of applications with which to use the metadata. The Save option saves the file locally for later use.
Indirect Import: When the Indirect Import option is selected, the metadata is displayed and may be copied and pasted as needed.
Comments
Comments must be written in English. Spam, offensive content, impersonation, and private information will not be permitted. If any comment is reported and identified as inappropriate content by OAE staff, the comment will be removed without notice. If you have any queries or need any help, please contact us at support@oaepublish.com.