Next Article in Journal
X-Ray Emissions from Hydrogen Rydberg Matter Detected Using Timepix3 CdTe Detector
Previous Article in Journal
Photocatalytic Conversion of β-O-4 Lignin Model Dimers: The Effect of Benzylic Ketones on Reaction Pathway
Previous Article in Special Issue
Zinc Indium Sulfide Materials for Photocatalytic Hydrogen Production via Water Splitting: A Short Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Facile Synthesis of the Single-Atom Decorated Cox-MoS2/RGO Catalysts by Thermal-Annealing Vacancy-Filling Strategy for Highly Efficient Hydrogen Evolution

1
School of Chemistry and Chemical Engineering, Jiangsu University, Zhenjiang 212013, China
2
School of Materials Science and Engineering, Jiangsu University, Zhenjiang 212013, China
*
Authors to whom correspondence should be addressed.
Catalysts 2025, 15(6), 524; https://doi.org/10.3390/catal15060524
Submission received: 6 May 2025 / Revised: 23 May 2025 / Accepted: 24 May 2025 / Published: 26 May 2025
(This article belongs to the Special Issue Recent Advances in Photo/Electrocatalytic Water Splitting)

Abstract

:
A “thermal-annealing vacancy-filling” synthesis strategy was developed to engineer cobalt single-atom catalysts (Co-MoS2/RGO) for exceptional hydrogen evolution reaction (HER) performance. By anchoring atomic Co onto Frenkel defect-engineered MoS2 nanosheets supported by reduced graphene oxide (RGO), we achieved simultaneous optimization of catalytic stability, electrical conductivity, and active site accessibility. The optimized Co3-MoS2/RGO hybrid demonstrates remarkable alkaline HER activity, requiring only 94.0 mV overpotential to achieve 10 mA cm−2 current density while maintaining excellent durability over extended operation. The atomically dispersed Co promoted HER kinetics through electronic structure modulation of MoS2 basal planes, creation of catalytic active centers, and defect-mediated synergies. The RGO further contributed to performance enhancement by preventing nanosheet aggregation, facilitating charge transfer, and exposing active sites. This defect engineering strategy provides a facile method for developing cost-effective, stable, and high-performance electrocatalysts for sustainable hydrogen production.

Graphical Abstract

1. Introduction

Green hydrogen, generated through renewable-powered water electrolysis with zero carbon emissions, has emerged as a pivotal eco-friendly alternative in sustainable energy solutions, particularly for decarbonizing industries such as steelmaking, ammonia synthesis, and long-haul transportation [1,2,3]. However, the sluggish kinetics of water decomposition, especially in alkaline conditions, present a significant bottleneck due to the additional energy required to overcome the high activation barrier for breaking H-OH bonds [4]. While precious metal-based catalysts, such as platinum (Pt) and Pt-based materials, have demonstrated high efficiency in water splitting, their high cost and limited availability hinder widespread practical applications [5,6]. To overcome these issues, researchers are increasingly focusing on alternative, cost-effective catalysts that offer high electrocatalytic performance [7].
Transition metal dichalcogenides (TMDs) have emerged as a highly investigated material class in electrocatalysis research, attributed to their cost-effectiveness, natural abundance, and superior catalytic performance in critical clean energy applications such as hydrogen evolution reactions (HER) [8,9,10]. Among TMDs, MoS2 has shown promising electrochemical stability and catalytic performance [11,12,13]. The active sites for electrocatalysis in MoS2 are typically found at the edge sites, where the coordination of atoms is unsaturated, whereas the basal planes remain catalytically inert due to their fully coordinated Mo-S bonds. As a result, pure MoS2 often exhibits suboptimal HER activity [14]. This limitation has spurred innovative strategies to activate the basal planes, including defect engineering [15], phase transition [16], surface modification [17], and hybridization with other materials [18], aimed at optimizing the electrocatalytic efficiency of MoS2 by strategically increasing the catalytic active surface sites.
Defect engineering, particularly the creation of point defects such as vacancies and interstitial atoms, has proven effective in enhancing the electrocatalytic properties of MoS2 [19,20,21]. For instance, the introduction of Frenkel defects in monolayer MoS2 has been shown to significantly improve HER performance, achieving a reduced overpotential of 164 mV at 10 mA cm−2 [22]. Moreover, doping MoS2 with heteroatoms such as Pd single atoms can activate inert atoms in the basal plane, further improving HER efficiency [23]. However, traditional single-atom synthesis methods are challenged by limitations such as low atomic loading density. Furthermore, the use of precious metal dopants often complicates synthesis and increases costs. Thus, there is growing interest in exploring non-precious metals for defect engineering in MoS2 as a more cost-effective and scalable strategy.
In this work, we propose a “thermal-annealing vacancy-filling” strategy to synthesize single-atomically dispersed Co-doped MoS2 catalysts. By creating Frenkel defects and filling Co atoms into MoS2 matrix, we achieved targeted augmentation of the catalytically active sites of MoS2 basal planes. To address nanoparticle aggregation, reduced graphene oxide (RGO) is employed as a support material, which simultaneously enhances charge transfer kinetics and surface dispersion stability [24]. Aberration-corrected scanning transmission electron microscopy (AC-STEM) coupled with synchrotron X-ray absorption spectroscopy (SXAS) confirmed the atomic dispersion of Co in MoS2 matrix. Electrochemical tests in alkaline electrolytes show that the HER performance of the obtained Co-MoS2/RGO catalyst is significantly improved.

2. Results and Discussion

Scheme 1 illustrates the key steps for synthesizing single-atom dispersed Cox-MoS2/RGO catalysts using the “thermal-annealing vacancy-filling” strategy. The process begins with the hydrothermal synthesis of MoS2 flower-like nanoparticles (NPs), followed by thermal annealing to induce Frenkel defects in the basal plane of the 2H-MoS2 sheets [22]. During thermal annealing under an argon atmosphere, Mo atoms migrate from their lattice positions, creating interstitial Mo atoms and Mo vacancies. Single-atomic Co atoms are then introduced to Frenkel defects and substitute the interstitial Mo atoms via the formation of S-Co-S bonds in the subsequent hydrothermal treatment in CoCl2 solution. The concentration of the Co precursor solution is adjusted to control the amount of Co loading. An additional annealing step further promotes the creation of Frenkel defects and enhances Mo vacancies. This approach not only activates previously inert in-plane atoms, transforming them into active sites, but also stabilizes the Frenkel defects, which are crucial for boosting the overall catalytic performance [22,23]. Finally, the Cox-MoS2 catalyst is supported on RGO to prevent aggregation of Co-MoS2 nanosheets, ensuring maximal exposure of active edges and basal planes. The resulting 3D structure of Cox-MoS2/RGO not only modulates the electronic and geometric structures but also prevents Co atoms from directly contacting acidic or basic solutions, thus protecting it from oxidation. Simultaneously, electron conductivity is maintained as electrons can efficiently travel through the MoS2 sheets, resulting in outstanding electrocatalytic performance for water splitting. The Co doping levels were controlled by varying the concentration of CoCl2 (0.25, 0.5, 0.75, and 1.0 mM) to yield Co1-MoS2, Co2-MoS2, Co3-MoS2, and Co4-MoS2, respectively. It should be pointed out that the same sample was synthesized more than three times, with no significant variations observed in the color, yield, or structural properties.
Figure 1A,B present representative scanning electron microscopy (SEM) images of MoS2 and Co3-MoS2 NPs, revealing a flower-like spherical structure with an average diameter of approximately 313 nm, showing that Co doping does not significantly alter the morphology. Transmission electron microscope (TEM) images in Figure 1C further confirm that the Co3-MoS2 NPs consist of interconnected thin sheets. These petal-like sheets not only provide a high specific surface area but also expose abundant MoS2 edges, which are typically the active sites for catalysis. Figure 1D shows the SEM image of Co3-MoS2/RGO NPs, where the wrinkled planar RGO supports the Co3-MoS2 NPs, preventing them from aggregating and restacking. No Co-related NPs or clusters are observed. To confirm the atomic dispersion of Co atoms, aberration-corrected high-angle annular dark-field scanning transmission electron microscopy (AC-HAADF-STEM) was employed. The AC-HAADF-STEM image in Figure 1E shows faintly colored Co atoms distributed across the MoS2 matrix; further, the corresponding intensity of the line scan confirms the single-atom dispersion of Co (Figure 1F). The atomic dispersion of Co was quantitatively analyzed using inductively coupled plasma (ICP), revealing a Co content of 3.21 wt% in Co3-MoS2. Table S1 in Supplementary Materials listed the weight percentages of the as-prepared Cox-MoS2 samples quantified by ICP analysis. When the Co loading increased to 4.80 wt% in Co4-MoS2, some Co clusters and aggregated particles were observed (Figure S1), indicating that excessive Co loading leads to aggregation. Thus, a Co loading of 3.21 wt% is optimal for achieving uniform Co atom dispersion. Figure 1G shows the energy dispersive X-ray spectrometer (EDS) mapping of Co3-MoS2/RGO, confirming the existence of Co and the homogeneous distribution of Mo, S, Co, and C elements across the Co3-MoS2/RGO. Additional EDS analysis (Figure S2 and Table S2) provides further details on the elemental composition of the as-prepared Cox-MoS2/RGO samples, confirming the expected molar percentages of Mo, S, Co, and C, which are consistent with the ICP analysis result.
Powered X-ray diffraction (XRD) patterns of the synthesized samples are presented in Figure 2A. The diffraction peaks of Cox-MoS2 are similar to those of pristine MoS2, indicating that Co doping does not significantly alter the crystalline structure. The main peaks correspond to the (100), (111), and (122) characteristic planes, which match the hexagonal MoS2 (2H-MoS2, JCPDS card no. 37-1492). Slight shifts in the peak positions, such as a left-shift of the (002) peak at 14.1° and a weakening of the (103) peak at 39.7°, suggest that Co doping introduces minor distortions to the MoS2 lattice, reducing crystallinity [25]. Importantly, no distinct diffraction peaks corresponding to cobalt containing phases are observed, indicating that Co element is present in an atomically dispersed form rather than as metallic clusters or compounds.
Raman spectroscopy was used to examine structural changes in the MoS2 and Cox-MoS2 samples. Figure 2B shows that MoS2 exhibits characteristic peaks at 375 cm−1 (E2g) and 402 cm−1 (A1g), confirming the existence of the 2H-MoS2 phase [26,27]. The Raman profiles of the Co-doped MoS2 samples remain largely unchanged, indicating that Co doping does not affect the crystalline structure of MoS2. Notably, two characteristic peaks centered at 1346 and 1587 cm−1 (Figure 2C), corresponding to RGO’s defect-induced (D band) and graphitic (G band) vibrational modes, are observed. With the increase of Co doping in the Cox-MoS2/RGO samples, the notably enhanced D/G intensity ratio (ID/IG) signifies increasing structural disorder within the graphene lattice, a critical feature that facilitates active site exposure through defect-mediated interfacial interactions while preserving charge transport pathways [21].
Figure 2D presents Fourier-transform infrared (FTIR) spectra of MoS2 and various Cox-MoS2/RGO samples. The FTIR spectrum of MoS2 shows a broad peak near 3435 cm−1, attributed to the stretching vibration of adsorbed water molecules. In addition, two peaks at 1621 and 1189 cm−1 correspond to the bending of adsorbed water and the asymmetric stretching of S=O bonds, respectively. These features suggest that water is present in the interlayer spaces of Cox-MoS2/RGO samples, which likely aids in its electrochemical activity [14].
X-ray photoelectron spectroscopy (XPS) was further employed to confirm the elemental composition and valence states of the as-synthesized Co3-MoS2/RGO. Figure 3A shows the XPS survey spectrum of the Co3-MoS2/RGO, confirming the presence of Mo, S, Co, C, and O elements across the sample. The high-resolution Mo 3d spectrum (Figure 3B) shows peaks corresponding to Mo4+ and Mo6+ states [28,29], while the high-resolution S 2p spectrum (Figure 3C) indicates the presence of S ions in MoS2. The high-resolution Co 2p spectrum (Figure 3D) shows peaks for Co ions and their associated satellite peaks, without an observable zero valence Co-Co peak [30], confirming the formation of Co single atoms within the MoS2 matrix. The high-resolution C 1s and O 1s spectra were shown in Figure S3, which indicates that a small amount of oxygen-containing groups are still retained on the surface of RGO, which is beneficial for the stable loading of Cox-MoS2 NPs.
The electronic configuration of Co species in Co3-MoS2/RGO catalyst was elucidated through SXAS spectroscopy. As depicted in Figure 4A, the normalized Co K-edge X-ray absorption near edge structure (XANES) profile of Co3-MoS2/RGO exhibits an absorption threshold energy of 7710.6 eV, positioned intermediately between Co foil (7705.4 eV) and CoO (7714.0 eV) reference standards. This characteristic pre-edge feature (1s→3d transition at 7700–7715 eV) demonstrates a 5.2 eV positive shift relative to Co foil while maintaining a 3.4 eV negative displacement compared to CoO, suggesting a mixed oxidation state regime spanning Co0 to Co2+. Figure 4B shows Fourier-transformed k2-weighted extended XAFS (EXAFS) spectra of Co3-MoS2/RGO in R-space that show notable differences compared to Co foil, CoO and Co3O4. A clear peak at ~1.64 Å in the R-space spectrum corresponds to the Co-S bond, while the absence of peaks at 2.19 Å (Co-Co) and 1.71 Å (Co-O) confirms that Co atoms are predominantly single-atomically within the MoS2 matrix via Co-S bonds. Wavelet transforms (WT) of the Co K-edge EXAFS (Figure 4D) further support this, showing a signal center at 1.64 Å in R-space, corresponding to the Co-S scattering path. Quantitative EXAFS fitting was performed to determine the Co-S coordination structure, with the best-fitting model shown in Figure 4C. The results reveal a Co-S coordination number of approximately 3.5 (Table S3), suggesting Co atoms occupy trigonal prismatic sites within the MoS2 lattice, consistent with theoretical predictions for highly active Co-MoS2 configurations [25].
The electrocatalytic HER performance of the as-synthesized Cox-MoS2/RGO samples was systematically evaluated through comprehensive electrochemical characterization. All measurements were conducted using a standard three-electrode system in 1 M KOH electrolyte, with a saturated calomel electrode (SCE) as reference, a graphite rod as a counter electrode, and a catalyst-modified glassy carbon electrode (GCE) as working electrode. As shown in Figure 5A, the linear sweep voltammetry (LSV) curves demonstrate that Co-doped MoS2/RGO hybrids exhibit significantly enhanced HER activity compared to pristine MoS2. Particularly, the optimized Co3-MoS2/RGO catalyst requires only 94.0 mV overpotential to achieve a current density of 10 mA cm−2, representing a 41% reduction compared to undoped MoS2 (160 mV) and approaching the commercial Pt/C catalyst (~45 mV) [31]. The Co3-MoS2/RGO displayed a competitive HER performance compared with the recently reported Co-doped MoS2 or MoS2/RGO hybrids (Table S4) [32,33,34,35,36,37,38,39]. This remarkable enhancement could be attributed to the synergistic effects of single-atom Co doping-induced electronic structure modulation and RGO’s superior charge transport properties [14,22]. Replicate LSV scans were conducted three times to confirm the consistency and accuracy in representing the corresponding LSV curves. Further kinetic analysis through Tafel plots (Figure 5B) reveals that Co3-MoS2/RGO exhibits the lowest Tafel slope of 63.6 mV dec−1, suggesting fast electron transfer kinetics during the HER process. Electrochemical active surface area (EASA) evaluation via cyclic voltammetry (CV) demonstrates that Co3-MoS2/RGO possesses the largest double-layer capacitance (Cdl = 166.4 mF cm−2) among the Cox-MoS2/RGO series (Figure 5C,D and Figure S4), indicating the high density of active sites for HER. The Co3-MoS2/RGO catalyst demonstrates the smallest overpotential and Tafel slope and the highest Cdl value, showing a faster electron transfer rate and suggesting that the rate-determining step is dominated by the Volmer-Heyrovsky pathway [40]. Notably, the Co3-MoS2/RGO catalyst demonstrates exceptional operational stability, retaining 88% of its initial activity after 3000 accelerated CV cycles (Figure 5E) and maintaining stable potential during 10 h chronopotentiometric testing (Figure 5F). This durability enhancement stems from RGO’s structural reinforcement and single-atom Co doping-induced stabilization of the MoS2 matrix [14,23]. The systematic improvement in both activity and stability positions Co3-MoS2/RGO as a promising non-precious metal catalyst for alkaline water electrolysis applications.
The improved HER performance of Cox-MoS2/RGO can be attributed to several synergistic mechanisms: Firstly, the doping of Co induces band structure reconfiguration in MoS2, narrowing the electronic bandgap and optimizing hydrogen adsorption free energy (ΔGH*) to thermoneutral hydrogen adsorption free energy [41]. These states enhance charge transfer kinetics and reduce the energy barrier for proton adsorption and H2 desorption; Secondly, the thermal-annealing vacancy-filling strategy can induce Frenkel-like defects (S vacancies and Mo interstitials), and then Co atoms replace Mo or occupy interstitial positions, which expose under-coordinated coordination sites, further increasing active site density. Thirdly, RGO’s high electrical conductivity mitigates the inherently poor conductivity of semiconducting 2H-MoS2, facilitates rapid electron transfer between Cox-MoS2/RGO and the electrode, lowering overpotential, and improving HER kinetics. Moreover, RGO can disperse the Cox-MoS2 NPs and prevent their restacking and Cox-MoS2 corrosion. This maximizes the exposure of coactive sites, increasing the density of catalytically active centers.

3. Experimental

3.1. Materials

Reagents including cobalt chloride hexahydrate (CoCl2·6H2O, 99%), ammonium molybdate tetrahydrate ((NH4)6Mo7O24·4H2O, 99%), thiourea (CS(NH2)2, 99%), and potassium hydroxide (KOH, 85%) were procured from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). Natural graphite powder with a particle size of 44 µm was supplied by Qingdao Zhongtian Graphite Co., Ltd. (Qingdao, China). A 5% Nafion perfluorinated resin solution was acquired from Alfa Aesar (Shanghai, China). All chemical reagents were utilized as received without undergoing additional purification. In all experimental procedures, Milli-Q water with a resistivity of 18.2 MΩ·cm−1 was used.

3.2. Synthesis of Flower-like Spherical MoS2

Flower-like MoS2 nanospheres were synthesized using a combined hydrothermal synthesis and thermal annealing approach. A 1:5 molar proportion of (NH4)6Mo7O24·4H2O and CS(NH2)2 was first dissolved in 40 mL of ultrapure water under continuous stirring. The resulting homogeneous solution was transferred into a 50 mL polytetrafluoroethylene-lined autoclave, which was then subjected to hydrothermal treatment at 200 °C for 36 h. Following natural cooling to room temperature, the reaction mixture underwent centrifugation, followed by repeated washing with ultrapure water. The collected solid was dried under vacuum at 60 °C to obtain as-synthesized MoS2. Subsequently, the MoS2 product was annealed at 350 °C for 2 h in an argon atmosphere to induce the formation of Frenkel defects.

3.3. Synthesis of Cox-MoS2

To synthesize Co-doped MoS2, the Frenkel defect MoS2 was dispersed in a 40 mL aqueous solution of cobalt chloride (0.75 mM) and sonicated for 30 min. The mixture was then subjected to hydrothermal treatment at 200 °C for 8 h. After cooling, the product was collected by centrifugation, washed, dried, and annealed at 350 °C under argon. The Co doping levels were controlled by varying the concentration of cobalt chloride (0.25, 0.5, 0.75, and 1.0 mM) to yield Co1-MoS2, Co2-MoS2, Co3-MoS2, and Co4-MoS2, respectively.

3.4. Synthesis of Cox-MoS2/RGO

GO was synthesized using potassium permanganate oxidation of natural graphite following Hummers’ method [42]. GO (10 mg) was dispersed in 15 mL of water to form a colloid. The above Cox-MoS2 powder was dispersed in 15 mL of water and then added to the GO colloid and sonicated, followed by magnetic stirring for 1 h. 1 mL of NaBH4 (0.15 M) was added to reduce the GO and form Cox-MoS2/RGO, which was then purified by centrifugation/wash cycles and dried under vacuum.

3.5. Characterization

The as-synthesized materials were analyzed using various techniques, including powder XRD (D8 Advanced, Bruker, Billerica, MA, USA), SEM (FEI NanoSEM450, Thermo Fisher Scientific, Hillsboro, OR, USA), TEM (JEM-2100, JEOL, Tokyo, Japan), HAADF-STEM (FEI Themis Z, Thermo Fisher Scientific, Hillsboro, OR, USA), FTIR (Nicolet Nexus 470, Thermo Fisher Scientific, Madison, WI, USA) spectroscopy, XPS (Thermo ESCALAB250, Thermo Fisher Scientific, Waltham, MA, USA), and ICP (VISTA-MPX, Agilent, Santa Clara, CA, USA) spectrometry.

3.6. Raman Measurements

Raman spectra were acquired using a Thermo Scientific DXR confocal microscope with a 532 nm diode-pumped excitation laser source (Thermo Fisher Scientific, Waltham, MA, USA). The excitation beam was focused onto the samples using an Olympus LMPlanFL 50× objective (0.50 NA, long-working-distance), yielding a spot size of approximately 2 µm. The laser power was maintained at 2 mW, and a grating with 1800 lines mm−1 (2 cm−1 spectral resolution) was employed for dispersion. Each spectrum was accumulated over three acquisitions (10 s integration time per scan).

3.7. Electrochemical Measurements

Electrochemical experiments were performed on a CHI 660E electrochemical workstation in 1 M KOH electrolyte (Sinopharm Chemical Reagent Co., Ltd., Shanghai, China). A glassy carbon electrode (GCE) loaded with the catalyst served as the working electrode, while a graphite rod and a saturated calomel electrode (SCE) functioned as the counter and reference electrodes, respectively. LSV and CV measurements were employed to assess HER activities.

4. Conclusions

In summary, we demonstrate a defect-engineering strategy, thermal-annealing vacancy-filling, to synthesize single-atomically dispersed Co-MoS2/RGO catalysts. This approach enables precise anchoring of Co atoms at Frenkel defect sites within the basal planes of 2H-MoS2, synergistically coupled with the conductive RGO. The optimized Co3-MoS2/RGO configuration achieves exceptional alkaline HER activity, requiring only 94.0 mV overpotential at 10 mA cm−2, rivaling state-of-the-art noble metal benchmarks. More importantly, this methodology offers a facile, scalable synthesis route to develop cost-competitive, high-performance electrocatalysts for sustainable hydrogen energy systems.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal15060524/s1, Figure S1: AC-HAADF-STEM image of the Co4-MoS2/RGO, Figure S2: EDS spectrum of the Co3-MoS2/RGO, Figure S3: High-resolution (A) C 1s and (B) O 1s XPS spectra of the Co3-MoS2/RGO, Figure S4: CV curves of the Co1-MoS2/RGO, Co2-MoS2/RGO, Co3-MoS2/RGO and Co4-MoS2/RGO with the scanning rate from 20 to 200 mV s−1; Table S1: Weight percentages of the as-prepared Cox-MoS2 samples quantified by ICP analysis, Table S2: Atomic percentages of the as-prepared Cox-MoS2/RGO samples quantified by EDS, Table S3: Parameters of EXAFS fits for the Co3-MoS2/RGO, Table S4: Comparison of the HER performance of recent reported Co-doped MoS2 or MoS2/RGO hybrids.

Author Contributions

J.Y.: Writing—original draft, Data curation. W.L.: Validation, Formal analysis. A.-A.A.: Validation, Formal analysis. X.L.: Validation, Data curation. J.N.: Validation, Formal analysis. S.W.: Validation, Conceptualization, Funding acquisition. X.F.: Writing—review and editing, Writing—original draft, Supervision. All authors have read and agreed to the published version of the manuscript.

Funding

The authors acknowledge the support from the National Natural Science Foundation of China (21776177) and in part from the Jiangsu Postdoctoral Research Foundation (2018K058C).

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Zhou, P.; Navid, I.A.; Ma, Y.; Xiao, Y.; Wang, P.; Ye, Z.; Zhou, B.; Mi, Z. Solar-to-hydrogen efficiency of more than 9% in photocatalytic water splitting. Nature 2023, 613, 66–70. [Google Scholar] [CrossRef] [PubMed]
  2. Ueckerdt, F.; Verpoort, P.C.; Anantharaman, R.; Bauer, C.; Beck, F.; Longden, T.; Roussanaly, S. On the cost competitiveness of blue and green hydrogen. Joule 2024, 8, 104–128. [Google Scholar] [CrossRef]
  3. Yu, H.; Díaz, A.; Lu, X.; Sun, B.; Ding, Y.; Koyama, M.; He, J.; Zhou, X.; Oudriss, A.; Feaugas, X.; et al. Hydrogen Embrittlement as a Conspicuous Material Challenge-Comprehensive Review and Future Directions. Chem. Rev. 2024, 124, 6271–6392. [Google Scholar] [CrossRef] [PubMed]
  4. Nishioka, S.; Osterloh, F.E.; Wang, X.; Mallouk, T.E.; Maeda, K. Photocatalytic water splitting. Nat. Rev. Method. Primers 2023, 3, 42. [Google Scholar] [CrossRef]
  5. Liu, J.; Li, Y.; Zhou, X.; Jiang, H.; Yang, H.; Li, C. Positively charged Pt-based cocatalysts: An orientation for achieving efficient photocatalytic water splitting. J. Mater. Chem. A 2020, 8, 17–26. [Google Scholar] [CrossRef]
  6. Zhou, Y.; Guo, W.; Xing, L.; Dong, Z.; Yang, Y.; Du, L.; Xie, X.; Ye, S. Keys to Unravel the Stability/Durability Issues of Platinum-Group-Metal Catalysts toward Oxygen Evolution Reaction for Acidic Water Splitting. ACS Cent. Sci. 2024, 10, 2006–2015. [Google Scholar] [CrossRef]
  7. Yang, R.; Fan, Y.; Zhang, Y.; Mei, L.; Zhu, R.; Qin, J.; Hu, J.; Chen, Z.; Ng, Y.N.; Voiry, D.; et al. 2D Transition Metal Dichalcogenides for Photocatalysis. Angew. Chem. Int. Ed. 2023, 62, 107519. [Google Scholar]
  8. Lu, Q.; Yu, Y.; Ma, Q.; Chen, B.; Zhang, H. 2D Transition-Metal-Dichalcogenide-Nanosheet-Based Composites for Photocatalytic and Electrocatalytic Hydrogen Evolution Reactions. Adv. Mater. 2015, 28, 1917–1933. [Google Scholar] [CrossRef]
  9. Wang, F.; Shifa, T.A.; Zhan, X.; Huang, Y.; Liu, K.; Cheng, Z.; Jiang, C.; He, J. Recent advances in transition-metal dichalcogenide based nanomaterials for water splitting. Nanoscale 2015, 7, 19764–19788. [Google Scholar] [CrossRef]
  10. He, Y.; Tang, P.; Hu, Z.; He, Q.; Zhu, C.; Wang, L.; Zeng, Q.; Golani, P.; Gao, G.; Fu, W.; et al. Engineering grain boundaries at the 2D limit for the hydrogen evolution reaction. Nat. Commun. 2020, 11, 57. [Google Scholar] [CrossRef]
  11. Ji, Q.; Yu, X.; Chen, L.; Yarley, O.P.N.; Zhou, C. Facile preparation of sugarcane bagasse-derived carbon supported MoS2 nanosheets for hydrogen evolution reaction. Ind. Crops Prod. 2021, 172, 114064. [Google Scholar] [CrossRef]
  12. Samy, O.; Zeng, S.; Birowosuto, M.D.; El Moutaouakil, A. A Review on MoS2 Properties, Synthesis, Sensing Applications and Challenges. Crystals 2021, 11, 355. [Google Scholar] [CrossRef]
  13. Liang, N.; Hu, X.; Li, W.; Wang, Y.; Guo, Z.; Huang, X.; Li, Z.; Zhang, X.; Zhang, J.; Xiao, J.; et al. A dual-signal fluorescent sensor based on MoS2 and CdTe quantum dots for tetracycline detection in milk. Food Chem. 2022, 378, 132076. [Google Scholar] [CrossRef] [PubMed]
  14. Shah, S.A.; Zhu, G.; Shen, X.; Kong, L.; Ji, Z.; Xu, K.; Zhou, H.; Zhu, J.; Song, P.; Song, C.; et al. Controllable Sandwiching of Reduced Graphene Oxide in Hierarchical Defect-Rich MoS2 Ultrathin Nanosheets with Expanded Interlayer Spacing for Electrocatalytic Hydrogen Evolution Reaction. Adv. Mater. Interfaces 2018, 5, 1801093. [Google Scholar] [CrossRef]
  15. He, Q.; Ye, N.; Han, L.; Tao, K. Sulfur Vacancy-Engineered Co3S4/MoS2-Interfaced Nanosheet Array for Enhanced Alkaline Overall Water Splitting. Inorg. Chem. 2023, 62, 21240–21246. [Google Scholar] [CrossRef] [PubMed]
  16. Zhao, Y.; Wei, S.; Wang, F.; Xu, L.; Liu, Y.; Lin, J.; Pan, K.; Pang, H. Hatted 1T/2H-Phase MoS2 on Ni3S2 Nanorods for Efficient Overall Water Splitting in Alkaline Media. Chem. Eur. J. 2020, 26, 2034–2040. [Google Scholar] [CrossRef]
  17. Liu, Y.; Hu, B.; Wu, S.; Wang, M.; Zhang, Z.; Cui, B.; He, L.; Du, M. Hierarchical nanocomposite electrocatalyst of bimetallic zeolitic imidazolate framework and MoS2 sheets for non-Pt methanol oxidation and water splitting. Appl. Catal. B 2019, 258, 117970. [Google Scholar] [CrossRef]
  18. Zhang, N.; Yang, Z.; Liu, W.; Zhang, F.; Yan, H. Novel Bifunctional Nitrogen Doped MoS2/COF-C4N Vertical Heterostructures for Electrocatalytic HER and OER. Catalysts 2023, 13, 90. [Google Scholar] [CrossRef]
  19. Lin, J.; Wang, P.; Wang, H.; Li, C.; Si, X.; Qi, J.; Cao, J.; Zhong, Z.; Fei, W.; Feng, J. Defect-Rich Heterogeneous MoS2/NiS2 Nanosheets Electrocatalysts for Efficient Overall Water Splitting. Adv. Sci. 2019, 6, 1900246. [Google Scholar] [CrossRef]
  20. Xie, J.; Qu, H.; Xin, J.; Zhang, X.; Cui, G.; Zhang, X.; Bao, J.; Tang, B.; Xie, Y. Defect-rich MoS2 nanowall catalyst for efficient hydrogen evolution reaction. Nano Res. 2017, 10, 1178–1188. [Google Scholar] [CrossRef]
  21. Tang, T.; Wang, Z.; Guan, J. A review of defect engineering in two-dimensional materials for electrocatalytic hydrogen evolution reaction. Chin. J. Catal. 2022, 43, 636–678. [Google Scholar] [CrossRef]
  22. Xu, J.; Shao, G.; Tang, X.; Lv, F.; Xiang, H.; Jing, C.; Liu, S.; Dai, S.; Li, Y.; Luo, J.; et al. Frenkel-defected monolayer MoS2 catalysts for efficient hydrogen evolution. Nat. Commun. 2022, 13, 2193. [Google Scholar] [CrossRef]
  23. Zhao, L.; Liang, S.; Zhang, L.; Huang, H.; Zhang, Q.; Ge, W.; Wang, S.; Tan, T.; Huang, L.; An, Q. Stabilizing and Activating Active Sites: 1T-MoS2 Supported Pd Single Atoms for Efficient Hydrogen Evolution Reaction. Small 2024, 20, 2401537. [Google Scholar] [CrossRef]
  24. Politano, G.G. Optimizing Graphene Oxide Film Quality: The Role of Solvent and Deposition Technique. C 2024, 10, 90. [Google Scholar] [CrossRef]
  25. Gong, J.; Zhang, Z.; Xi, S.; Wang, W.; Lu, J.; Chen, P. Graphene quantum dot enabled interlayer spacing and electronic structure regulation of single-atom doped MoS2 for efficient alkaline hydrogen evolution. Chem. Eng. J. 2023, 451, 138951. [Google Scholar] [CrossRef]
  26. Koudakan, P.A.; Wei, C.; Mosallanezhad, A.; Liu, B.; Fang, Y.; Hao, X.; Qian, Y.; Wang, G. Constructing Reactive Micro-Environment in Basal Plane of MoS2 for pH-Universal Hydrogen Evolution Catalysis. Small 2022, 18, 2107974. [Google Scholar] [CrossRef]
  27. Politano, G.G.; Castriota, M.; Santo, M.P.D.; Pipita, M.M.; Desiderio, G.; Vena, C.; Versace, C. Variable angle spectroscopic ellipsometry characterization of spin-coated MoS2 films. Vacuum 2021, 189, 110232. [Google Scholar] [CrossRef]
  28. Liang, J.; Li, H.; Chen, L.; Ren, M.; Fakayode, O.A.; Han, J.; Zhou, C. Efficient hydrogen evolution reaction performance using lignin-assisted chestnut shell carbon-loaded molybdenum disulfide. Ind. Crops Prod. 2022, 193, 116214. [Google Scholar] [CrossRef]
  29. Wang, X.; Pan, Y.; Wang, X.; Guo, Y.; Ni, C.; Wu, J.; Hao, C. High performance hybrid supercapacitors assembled with multi-cavity nickel cobalt sulfide hollow microspheres as cathode and porous typha-derived carbon as anode. Ind. Crops Prod. 2022, 189, 115863. [Google Scholar] [CrossRef]
  30. Wu, Q.; Zhong, Y.; Chen, R.; Ling, G.; Wang, X. Cu-Ag-C@Ni3S4 with core shell structure and rose derived carbon electrode materials: An environmentally friendly supercapacitor with high energy and power density. Ind. Crops Prod. 2024, 222, 119676. [Google Scholar] [CrossRef]
  31. Pearson, A.; O’Mullane, A.P. A simple approach to improve the electrocatalytic properties of commercial Pt/C. Chem. Commun. 2015, 51, 11297–11300. [Google Scholar] [CrossRef] [PubMed]
  32. Zhou, F.; Zhang, X.; Sa, R.; Zhang, S.; Wen, Z.; Wang, R. The Electrochemical Overall Water Splitting Promoted by MoS2 in Coupled Nickel–Iron (Oxy)Hydride/Molybdenum Sulfide/Graphene Composite. Chem. Eng. J. 2020, 397, 125454. [Google Scholar] [CrossRef]
  33. Zhang, C.; Xie, Y.; Liu, J.; Cao, F.; Cong, H.; Li, H. 1D Core–Shell MOFs Derived CoP Nanoparticles-Embedded N-Doped Porous Carbon Nanotubes Anchored with MoS2 Nanosheets as Efficient Bifunctional Electrocatalysts. Chem. Eng. J. 2021, 419, 129977. [Google Scholar] [CrossRef]
  34. Chen, Y.; Meng, G.; Yang, T.; Chen, C.; Chang, Z.; Kong, F.; Tian, H.; Cui, X.; Hou, X.; Shi, J. Interfacial engineering of Co-doped 1T-MoS2 coupled with V2C MXene for efficient electrocatalytic hydrogen evolution. Chem. Eng. J. 2022, 450, 138157. [Google Scholar] [CrossRef]
  35. Kim, M.; Seok, H.; Clament Sagaya Selvam, N.; Cho, J.; Choi, G.H.; Nam, M.G.; Kang, S.; Kim, T.; Yoo, P.J. Kirkendall Effect Induced Bifunctional Hybrid Electrocatalyst (Co9S8@MoS2/N-Doped Hollow Carbon) for High Performance Overall Water Splitting. J. Power Sources 2021, 493, 229688. [Google Scholar] [CrossRef]
  36. Ma, W.; Li, W.; Zhang, H.; Wang, Y. N-doped carbon wrapped CoFe alloy nanoparticles with MoS2 nanosheets as electrocatalyst for hydrogen and oxygen evolution reactions. Int. J. Hydrogen Energy 2023, 48, 22032–22043. [Google Scholar] [CrossRef]
  37. Li, D.; Fan, S.; Li, J.; Li, W.; Zhuang, Y.; Li, Y.; Shan, L.; Dong, L.; Yao, J. Preparation and characterization of bifunctional 1T-2H MoS2-Sv/CuS catalyst for electrocatalytic hydrogen and oxygen evolution reaction. Sep. Purif. Technol. 2025, 352, 128176. [Google Scholar] [CrossRef]
  38. Liu, Q.; Xue, Z.; Jia, B.; Liu, Q.; Liu, K.; Lin, Y.; Liu, M.; Li, Y.; Li, G. Hierarchical Nanorods of MoS2/MoP Heterojunction for Efficient Electrocatalytic Hydrogen Evolution Reaction. Small 2020, 16, 2002482. [Google Scholar] [CrossRef]
  39. Luo, M.; Liu, S.; Zhu, W.; Ye, G.; Wang, J.; He, Z. An Electrodeposited MoS2-MoO3-x/Ni3S2 Heterostructure Electrocatalyst for Efficient Alkaline Hydrogen Evolution. Chem. Eng. J. 2022, 428, 131055. [Google Scholar] [CrossRef]
  40. Weng, Z.; Zhu, J.; Lu, L.; Ma, Y.; Cai, J. Regulation of the electric double-layer capacitance of MoS2/ionic liquid by carbon modification. J. Appl. Electrochem. 2023, 53, 689–703. [Google Scholar] [CrossRef]
  41. Zhou, Y.; Gao, G.; Li, Y.; Chu, W.; Wang, L. Transition-metal single atoms in nitrogen-doped graphenes as efficient active centers for water splitting: A theoretical study. Phys. Chem. Chem. Phys. 2019, 21, 3024–3032. [Google Scholar] [CrossRef]
  42. Hummers, W.S.; Offeman, R.E. Preparation of Graphitic Oxide. J. Am. Chem. Soc. 1958, 80, 1339. [Google Scholar] [CrossRef]
Scheme 1. Schematic illustration of the preparation process of Cox-MoS2/RGO catalysts.
Scheme 1. Schematic illustration of the preparation process of Cox-MoS2/RGO catalysts.
Catalysts 15 00524 sch001
Figure 1. SEM images of the (A) MoS2 and (B) Co3-MoS2 NPs. (C) TEM image of an individual Co3-MoS2 NPs. (D) SEM image of the Co3-MoS2/RGO NPs. (E) AC-HAADF-STEM image of the Co3-MoS2/RGO NPs. Yellow circles indicate Co atoms. (F) corresponding linear intensity profile of the blue rectangle in panel (D). (G) EDS elemental mapping of the Co3-MoS2/RGO NPs.
Figure 1. SEM images of the (A) MoS2 and (B) Co3-MoS2 NPs. (C) TEM image of an individual Co3-MoS2 NPs. (D) SEM image of the Co3-MoS2/RGO NPs. (E) AC-HAADF-STEM image of the Co3-MoS2/RGO NPs. Yellow circles indicate Co atoms. (F) corresponding linear intensity profile of the blue rectangle in panel (D). (G) EDS elemental mapping of the Co3-MoS2/RGO NPs.
Catalysts 15 00524 g001
Figure 2. (A) Powered XRD patterns and (B) Raman spectra of the synthesized MoS2, Co1-MoS2, Co2-MoS2, Co3-MoS2, and Co4-MoS2 NPs. (C) Raman spectra and (D) FTIR spectra of the Co1-MoS2/RGO, Co2-MoS2/RGO, Co3-MoS2/RGO, and Co4-MoS2/RGO NPs.
Figure 2. (A) Powered XRD patterns and (B) Raman spectra of the synthesized MoS2, Co1-MoS2, Co2-MoS2, Co3-MoS2, and Co4-MoS2 NPs. (C) Raman spectra and (D) FTIR spectra of the Co1-MoS2/RGO, Co2-MoS2/RGO, Co3-MoS2/RGO, and Co4-MoS2/RGO NPs.
Catalysts 15 00524 g002
Figure 3. (A) XPS survey spectrum of the Co3-MoS2/RGO NPs and corresponding high-resolution spectra of (B) Mo 3d, (C) S 2p, and (D) Co 2p.
Figure 3. (A) XPS survey spectrum of the Co3-MoS2/RGO NPs and corresponding high-resolution spectra of (B) Mo 3d, (C) S 2p, and (D) Co 2p.
Catalysts 15 00524 g003
Figure 4. (A) Co K-edge XANES spectra and (B) k2-weighted Fourier transform EXAFS spectra of Co3-MoS2/RGO along with reference samples (Co foil, CoO, and Co3O4) at R space, and corresponding (C) EXAFS fitting curve. The schematic structural model of Co-MoS2 is also presented. Mo, navy blue, Co, purple, S, yellow. (D) Wavelet transforms of the Co K-edge EXAFS signals.
Figure 4. (A) Co K-edge XANES spectra and (B) k2-weighted Fourier transform EXAFS spectra of Co3-MoS2/RGO along with reference samples (Co foil, CoO, and Co3O4) at R space, and corresponding (C) EXAFS fitting curve. The schematic structural model of Co-MoS2 is also presented. Mo, navy blue, Co, purple, S, yellow. (D) Wavelet transforms of the Co K-edge EXAFS signals.
Catalysts 15 00524 g004
Figure 5. (A) LSV curves and (B) corresponding Tafel slopes for the Co3-MoS2/RGO electrodes. (C) CV curves of Co3-MoS2/RGO at varying scan rates, with their derived (D) double-layer capacitance curves. (E) LSV curves of Co3-MoS2/RGO before and after 3000 CV cycles. (F) Chronopotentiometric stability test of Co3-MoS2/RGO at a fixed potential of −94 mV.
Figure 5. (A) LSV curves and (B) corresponding Tafel slopes for the Co3-MoS2/RGO electrodes. (C) CV curves of Co3-MoS2/RGO at varying scan rates, with their derived (D) double-layer capacitance curves. (E) LSV curves of Co3-MoS2/RGO before and after 3000 CV cycles. (F) Chronopotentiometric stability test of Co3-MoS2/RGO at a fixed potential of −94 mV.
Catalysts 15 00524 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Yang, J.; Li, W.; Aimeti, A.-A.; Liu, X.; Nie, J.; Wang, S.; Fu, X. Facile Synthesis of the Single-Atom Decorated Cox-MoS2/RGO Catalysts by Thermal-Annealing Vacancy-Filling Strategy for Highly Efficient Hydrogen Evolution. Catalysts 2025, 15, 524. https://doi.org/10.3390/catal15060524

AMA Style

Yang J, Li W, Aimeti A-A, Liu X, Nie J, Wang S, Fu X. Facile Synthesis of the Single-Atom Decorated Cox-MoS2/RGO Catalysts by Thermal-Annealing Vacancy-Filling Strategy for Highly Efficient Hydrogen Evolution. Catalysts. 2025; 15(6):524. https://doi.org/10.3390/catal15060524

Chicago/Turabian Style

Yang, Jiang, Wentao Li, Abdul-Aziz Aimeti, Xinyu Liu, Jiaqi Nie, Shuang Wang, and Xiaoqi Fu. 2025. "Facile Synthesis of the Single-Atom Decorated Cox-MoS2/RGO Catalysts by Thermal-Annealing Vacancy-Filling Strategy for Highly Efficient Hydrogen Evolution" Catalysts 15, no. 6: 524. https://doi.org/10.3390/catal15060524

APA Style

Yang, J., Li, W., Aimeti, A.-A., Liu, X., Nie, J., Wang, S., & Fu, X. (2025). Facile Synthesis of the Single-Atom Decorated Cox-MoS2/RGO Catalysts by Thermal-Annealing Vacancy-Filling Strategy for Highly Efficient Hydrogen Evolution. Catalysts, 15(6), 524. https://doi.org/10.3390/catal15060524

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop