Next Article in Journal
Catalytic Methods for Producing Fine and Bulk Chemicals and Biomaterials from Biomass
Previous Article in Journal
Supported TiO2 Photocatalysis of Spiked Contaminants in Water and Municipal Wastewater
Previous Article in Special Issue
Integrating Density Functional Theory Calculations and Machine Learning to Identify Conduction Band Minimum as a Descriptor for High-Efficiency Hydrogen Evolution Reaction Catalysts in Transition Metal Dichalcogenides
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Growth of Two-Dimensional Edge-Rich Screwed WS2 with High Active Site Density for Accelerated Hydrogen Evolution

1
Hebei Key Laboratory of Optic-Electronic Information and Materials, College of Physics Science and Technology, Hebei University, Baoding 071002, China
2
College of Life Science, Hebei University, Baoding 071002, China
3
College of Advanced Interdisciplinary Studies & Hunan Provincial Key Laboratory of Novel Nano-Optoelectronic Information Materials and Devices, National University of Defense Technology, Changsha 410073, China
*
Authors to whom correspondence should be addressed.
Catalysts 2025, 15(5), 496; https://doi.org/10.3390/catal15050496
Submission received: 10 April 2025 / Revised: 14 May 2025 / Accepted: 16 May 2025 / Published: 20 May 2025
(This article belongs to the Special Issue Two-Dimensional (2D) Materials in Catalysis)

Abstract

:
Two-dimensional transition metal dichalcogenides have attracted considerable attention in electrocatalytic hydrogen evolution due to their unique layered structures and tunable electronic properties. However, prior research has predominantly focused on the intrinsic catalytic activity of planar few-layer structures, which offer limited exposure of edge-active sites due to their restricted two-dimensional geometry. Moreover, van der Waals interactions between layers impose substantial barriers to electron transport, significantly hindering charge transfer efficiency. To overcome these limitations, this study presents the innovative synthesis of high-quality single-screw WS2 with a 5° dislocation angle via physical vapor deposition. Second harmonic generation measurements revealed a pronounced asymmetric polarization response, while the selected area electron diffractionand atomic force microscopy elucidated the material’s distinctive screwed dislocation configuration. In contrast to planar monolayer WS2, the conical/screw-structured WS2—formed through screw-dislocation-mediated growth—exhibits a higher density of exposed edge-active catalytic sites and enhanced electron transport capabilities. Electrochemical performance tests revealed that in an alkaline medium, the screwed WS2 nanosheets exhibited an overpotential of 310 mV at a current density of −10 mA/cm2, with a Tafel slope of 204 mV/dec. Additionally, under a current density of 18 mA/cm2, the screwed WS2 can sustain this current density for at least 30 h. These findings offer valuable insights into the design of low-cost, high-efficiency, non-precious metal catalysts for hydrogen evolution reactions.

Graphical Abstract

1. Introduction

Energy serves as the driving force behind the advancement of modern civilization, playing a pivotal role in human production and societal development. At present, the global energy supply system remains heavily reliant on non-renewable fossil fuels. However, the excessive consumption of these resources has intensified environmental pressures and raised concerns about resource depletion, presenting humanity with the dual challenges of environmental pollution and energy security.Among renewable energy technologies, hydrogen production via water electrolysis has garnered significant attention as a promising alternative to fossil fuels. The key to the widespread adoption of this technology lies in the development of efficient catalysts capable of lowering the reaction energy barrier. In this context, the design of electrode materials that combine high catalytic performance with cost-effectiveness has become a central focus of research [1,2,3,4]. Traditional noble metal catalysts, such as platinum (Pt), offer excellent catalytic activity but suffer from high costs and limited availability. This has driven the exploration of low-cost substitutes. Two-dimensional transition metal dichalcogenides (2D-TMDs)—including materials such as MoS2, WS2, and CrSe2—have emerged as promising alternatives due to their tunable electronic structures, high surface-to-volume ratios, and abundant active sites [5,6,7,8,9]. Theoretical studies have revealed that the Gibbs free energy values for hydrogen evolution in 2D-TMDs are comparable to those of platinum-based catalysts, indicating their strong potential to replace noble metals [10,11,12,13]. For instance, monolayer or multilayer homogeneous stacked transition metal dichalcogenides exhibit hydrogen evolution reaction catalytic performance comparable to that of platinum, despite relying solely on their limited edge-active catalytic sites [10,14,15,16,17,18]. Despite this potential, these semiconductor materials face two major limitations in practical applications: (1) the limited exposure of active edge sites restricts overall catalytic efficiency, and (2) their inherently low electrical conductivity hampers effective charge transport [19,20,21,22,23].
To overcome these challenges, researchers have pursued strategies such as increasing the density of active sites, enhancing the intrinsic activity of catalytic centers, and improving charge transport properties through structural engineering and performance optimization [20,21,24,25,26,27,28,29,30,31,32,33]. One particularly promising structural configuration is the “screwed” growth architecture observed in WS2 during the early stages of formation. Unlike conventional planar epitaxial growth, this unique structure—driven by screw dislocations—results in vertically stacked, spiral architectures that expose a significantly higher number of catalytically active sites. Density functional theory (DFT) calculations combined with experimental measurements in acidic electrolytes demonstrate that introducing dislocations, interlayer twisting, and stacking faults into screwed WS2 architectures substantially increases the density of edge sites and sulfur vacancies, thereby enhancing catalytic activity [28,29,30,31,32]. These sites often contain unsaturated sulfur coordination, which facilitates hydrogen adsorption and makes them efficient reactive centers [33,34,35,36,37,38,39,40]. Furthermore, the local lattice strain induced by the screwed architecture can modulate the electronic band structure of WS2, bringing the Gibbs free energy of hydrogen adsorption (ΔGH*) closer to the ideal value of zero. This reduces the reaction energy barrier, promotes proton reduction to molecular hydrogen, and enhances the overall HER efficiency in water electrolysis [39,40,41,42]. Additionally, the interlayer coupling and defect engineering within the screwed structure can enhance electrical conductivity, expedite charge transfer, and reduce electrochemical polarization [43,44,45,46,47,48]. The architecture also improves mechanical strength, helping to mitigate structural degradation during prolonged catalytic operations [41,42,43,44]. Given these advantageous physical and chemical properties, exploring the catalytic performance of screwed WS2 structures is of considerable scientific and technological interest.
In this study, high-quality screwed WS2 structures were successfully synthesized via bidirectional gas-flow physical vapor deposition. Atomic force microscopy (AFM) confirmed the characteristic screwed morphology of the resulting WS2. Selected area electron diffraction (SAED) and second harmonic generation (SHG) analyses revealed the presence of interlayer dislocations and structural asymmetry within the screwed WS2, highlighting its unique internal architecture. Unlike previous studies that typically evaluated the hydrogen evolution reaction (HER) performance of WS2 in acidic media, this work employs an alkaline electrolyte as the catalytic environment. In acidic conditions, HER primarily proceeds through the reduction of H⁺ ions, which can result in the deactivation of edge sites due to excessive proton adsorption. In contrast, alkaline media allow for surface charge modulation, mitigating this issue and preserving catalytic activity. Thanks to its abundant edge-active sites and interlayer distortion-induced defects, screwed WS2 exhibits significantly enhanced catalytic performance and electrical conductivity compared to conventional monolayer WS2. These improvements address the common limitations of low active site density and poor conductivity observed in typical WS2 catalysts. This advancement offers a promising strategy for the development of highly efficient, low-cost, non-precious metal catalysts for hydrogen evolution.

2. Results and Discussion

2.1. Growth Mechanism and Characterization

A schematic diagram illustrating the preparation process is shown in Figure S1. The simulated growth process of monolayer WS2 is depicted in Figure 1a. During the synthesis of monolayer WS2, growth initiates from a single nucleation site and proceeds outward in a lateral, planar manner—commonly referred to as in-plane growth. In contrast, the formation of screwed WS2 follows a screw dislocation-driven vertical growth mechanism, as shown in the dynamic evolution illustrated in Figure 1b. When two WS2 crystalline domains intersect, lattice distortion occurs at their interface, leading to the formation of a localized protrusion. At the edges of these protrusions, unsaturated sulfur atoms with dangling bonds accumulate, serving as catalytically active sites. By optimizing the growth conditions—specifically, employing low-temperature deposition and reducing precursor concentration—in-plane growth is effectively suppressed, while dislocation-driven vertical (screwed) growth is promoted. Optical microscopy images of monolayer WS2 and screwed WS2 nanostructures are shown in Figure 1c,d, respectively. Growth terminates once the precursor materials are depleted. X-ray photoelectron spectroscopy (XPS) was used to confirm the chemical composition of WS2, with results presented in Figure S2. The XPS spectra exhibit peaks at 32.9 eV and 35.1 eV, corresponding to the W 4f7/2 and W 4f5/2 states of tungsten, respectively. Peaks at 162.2 eV and 163.5 eV correspond to the S 2p3/2 and S 2p1/2 states of sulfur [49,50]. Raman spectroscopy was also conducted for both screwed and monolayer WS2 samples, as shown in Figure S3. The spectra exhibit characteristic peaks at ~350 cm−1 and ~420 cm−1, corresponding to the E12g and A1g phonon modes of WS2, respectively—confirming the successful synthesis of WS2 [51,52]. Atomic force microscopy (AFM) was employed to investigate the topography of the screwed WS2 structures. As shown in Figure 1e,f, the AFM images clearly reveal the single-screw morphology. Each screw layer is smooth and well-defined, with a measured interlayer thickness of approximately 0.56 nm [53,54]. For comparison, the AFM-measured thickness of monolayer WS2 is approximately 0.66 nm (Figure S4), consistent with known monolayer thickness values. Furthermore, the AFM profile in Figure 1g shows that the angle between the origin point of the screw structure and the base layer is approximately 5°. Selected area electron diffraction (SAED), shown in Figure 1h, reveals the lattice distortion associated with screw dislocations. The diffraction pattern obtained along the (100) crystallographic zone axis displays a shift of approximately 5°, aligning with the model of continuous rotational displacement of crystal planes induced by screw dislocations.

2.2. Optical Properties of Screwed Nanomaterials

SHG, a nonlinear optical phenomenon highly sensitive to crystalline orientation, serves as a powerful tool for probing the symmetry of two-dimensional layered materials. When a material lacks spatial inversion symmetry, its second-order nonlinear optical susceptibility becomes non-zero, leading to a pronounced SHG response [55,56]. Monolayer WS2 is known to exhibit a non-centrosymmetric crystal structure, as reported in previous studies [57,58]. AFM topography reveals that screwed WS2 possesses a structure resembling 3R-type stacking, which can be approximated as an “AA”-type stacking configuration. To further confirm the non-centrosymmetric structure induced by interlayer dislocations in screwed WS2, SHG characterization was performed. Systematic SHG measurements were conducted to probe the symmetry properties of the material, and the results are shown in Figure 2a. In the edge regions of the screwed steps—where non-centrosymmetry is more pronounced—the SHG signal intensity is approximately ten times higher than that of planar monolayer WS2, confirming a significant local symmetry breaking caused by the screw-dislocation-induced stacking. As shown in Figure 2b, the SHG intensity of screwed WS2 increases with rising laser power, further supporting the presence of a non-centrosymmetric structure. The extracted SHG intensities at various laser powers were plotted on a logarithmic scale (Figure 2c), yielding a slope of 2.07—closely matching the theoretical expectation of 2 for a second-order nonlinear optical process. Polarization-resolved SHG measurements were also carried out to investigate anisotropic structural features. As illustrated in Figure 2d, the SHG response of screwed WS2 displays a periodic modulation with a polarization angle offset of approximately 5°, consistent with the interlayer rotational angle induced by screw dislocations. Monolayer WS2 typically adopts a hexagonal 2H phase, with a sandwich-like S–W–S atomic configuration that lacks a center of inversion symmetry. This inherent asymmetry allows for the emergence of SHG signals. Additionally, the in-plane anisotropy of monolayer WS2 causes the SHG intensity to exhibit periodic modulation as a function of the incident light’s polarization angle (Figure S5). Comparative SHG measurements between screwed WS2 and monolayer WS2 reveal that the SHG intensity of screwed WS2 is approximately an order of magnitude higher. This enhancement is attributed to a higher density of interlayer defects and symmetry-breaking features, which are beneficial for catalytic performance. The pronounced SHG response of screwed WS2 thus correlates with its potential for significantly improved hydrogen evolution activity.

2.3. Electrocatalytic Testing

To evaluate the electrocatalytic performance of screwed WS2 and monolayer WS2, devices were fabricated by transferring pre-patterned electrodes onto the WS2 surfaces. Compared with conventional metal deposition methods such as magnetron sputtering or thermal evaporation, this transfer technique effectively mitigates interfacial issues, including Fermi level pinning and parasitic interfacial capacitance, thereby preserving the intrinsic catalytic and electrical characteristics of the WS2 nanostructures. For the fabrication of screwed WS2 device, a mechanically exfoliated few-layered mica flake was precisely transferred onto the central-to-edge region of the screwed WS2, serving as a dielectric spacer. This spatial configuration isolates the central and edge regions of WS2, allowing direct and reliable measurements of charge transport efficiency from the core to the periphery. Such a setup provides valuable insights into the anisotropic electron transport behavior of the screwed structure. The electronic transport properties of monolayer and screwed WS2 were measured, and the results are shown in Figure 3a,b, respectively. The corresponding optical image of the fabricated device is provided in the lower-right corner. Under an applied voltage sweep from −1 V to 1 V, the maximum current for monolayer WS2 reaches 0.9 nA, whereas the screwed WS2 achieves a significantly higher current of 5400 nA. This marked difference indicates that the screwed WS2 possesses superior electron transport efficiency, facilitating faster charge transfer and enhanced hydrogen evolution performance. For electrochemical testing, both WS2 samples (grown on SiO2/Si substrates) were transferred onto copper sheets via a wet-transfer method. A standard three-electrode configuration was employed: the WS2 samples served as the working electrodes, a mercury/mercuric oxide (Hg/HgO) electrode was used as the reference, and a graphene rod functioned as the counter electrode. Figure 3c,d present the linear sweep voltammetry (LSV) curves and the corresponding Tafel slopes for both WS2 morphologies. At a current density of −10 mA cm−2, the overpotential of screwed WS2 is 310 mV with a Tafel slope of 204 mV dec−1, while monolayer WS2 exhibits a higher overpotential of 390 mV and a Tafel slope of 225 mV dec−1. These results clearly demonstrate the superior HER performance of the screwed WS2 architecture. Electrochemical impedance spectroscopy (EIS) was further employed to assess the charge transfer kinetics of the two morphologies. As shown in Figure 3e, the charge transfer resistance (Rct) for screwed WS2 is 72 Ω, compared to 92 Ω for monolayer WS2, confirming improved interfacial charge transport in the screwed structure. The ECSA is a key metric for evaluating the number of active sites on a catalyst’s surface, and it is directly proportional to the electrochemical double-layer capacitance (ECSA = Cdl/Cs). In this work, the specific capacitance value (Cs) was taken as 40 μF cm−2. Cyclic voltammetry at different scan rates (5–25 mV s−1) was employed to determine the Cdl value by fitting and analyzing the relationship between current density and scan rate. The test results are shown in Figure S6. From the data, it can be observed that the screwed WS2 and copper substrate exhibit a Cdl of 403.8 mF cm−2, corresponding to an ECSA of 10,095 cm2, whereas the monolayer WS2 and copper substrate show a Cdl of 313.2 mF cm−2, yielding an ECSA of 7830 cm2. The results clearly demonstrate that the ECSA of screwed WS2 is significantly higher than that of monolayer WS2. To investigate the structural durability and catalytic stability, long-term performance retention tests were conducted. As shown in Figure 3f, the chronoamperometric experiments revealed that at a fixed overpotential, the current densities of screwed WS2 and monolayer WS2 remained stable at 18 mA cm−2 and 10.3 mA cm−2, respectively, for 30 h without degradation.

2.4. Electrocatalytic Mechanism Analysis

The electrocatalytic process for hydrogen evolution begins with the electrochemical adsorption step, which follows the Volmer reaction. In this step, hydrogen ions (protons) react with electrons on the catalyst surface, resulting in the adsorption of a hydrogen atom (H*) on the material surface. This corresponds to the following reaction (1). As the hydrogen content on the surface increases, the Tafel desorption reaction occurs (2), where two adjacent adsorbed hydrogen intermediates combine to form a hydrogen molecule. The mechanistic simulation diagrams for hydrogen evolution catalysis in alkaline solution (KOH) for screwed WS2 and monolayer WS2 are presented in Figure 4a,b, respectively.
H2O + e + *→H* + OH
2H*→H2
WS2 with a screwed structure possesses a continuous network of dislocations, providing a high density of surface active sites. This unique structural feature endows the material with exceptional catalytic properties. Compared to conventional two-dimensional materials, the screwed WS2 exhibits superior catalytic kinetics in the hydrogen evolution reaction (HER). Specifically, it demonstrates a smaller Tafel slope and a lower onset potential, both of which are indicators of enhanced electrocatalytic efficiency. Studies have shown that the current intensity generated by the screwed WS2 structure is approximately four orders of magnitude higher than that of the monolayer system. This significant enhancement is attributed to the fact that the screwed configuration allows the current to flow along the screw-dislocation edge, bypassing the need to overcome interlayer potential barriers. This geometry accelerates electron movement and optimizes the hydrogen evolution performance. In addition, lattice distortion at the interface of the screwed structure creates a localized bulging region, where unsaturated sulfur (S) atoms are suspended. These unsaturated sulfur atoms improve the adsorption of hydrogen atoms, facilitating the formation of H2 molecules when sufficient hydrogen is accumulated on the material surface. This characteristic further boosts the catalytic hydrogen evolution ability of the material. Moreover, the screwed structure not only enhances the catalytic activity but also improves mechanical strength, reducing the likelihood of structural collapse during the catalytic process. This ensures a more reliable and stable hydrogen evolution process, making screwed WS2 an effective and durable catalyst for hydrogen production.

3. Experimental Section

3.1. Synthesis of Screwed WS2 Nanosheets

Screwed WS2 and monolayer WS2 nanosheets with atomic thickness were synthesized on SiO2/Si substrates using the physical vapor deposition (PVD) method. A quartz tube (1 m length, 2.2 cm inner diameter) was placed in a horizontal high-temperature tube furnace (KSY controllable silicon temperature tube furnace, Shenyang Energy-saving Electric Furnace Factory, Shengyang, China). Using an electronic balance, 1 g of WS2 (99.8%, Alfa Aesar, Haverhill, MA, USA) solid powder was weighed and placed in a ceramic boat. The ceramic boat was positioned in the central high-temperature zone of the tube furnace. A SiO2/Si substrate (2 cm in length) was placed in a half-cut ceramic boat and pushed to the downstream deposition zone of the furnace using a small cart. The quartz tube was sealed with a rubber stopper. The purpose of the small cart was to immediately remove the substrate from the deposition zone after the temperature rise, allowing the WS2 to grow and then cool abruptly, which enhanced the crystallinity of the WS2 nanosheets. Additionally, extracting the substrate prevented re-nucleation of already grown nanosheets. For monolayer WS2 nanosheets, the target heating temperature was set to 1100 °C, with an argon gas flow rate of 60 sccm (99.999%, Baoding Jinglian Gas Co., Ltd., Baoding, China). Reverse annealing was performed for 60 min, followed by forward annealing at 60 sccm for 10 min to prepare the monolayer WS2 nanosheets. For screwed WS2 nanosheets, the target heating temperature was set to 1070 °C, with reverse annealing for 10 min, followed by forward annealing at 60 sccm for 10 min to prepare the screwed WS2 nanosheets.

3.2. Catalytic Sample Preparation

The WS2 nanosheets (screwed/monolayer form) prepared on SiO2/Si substrates were transferred onto copper sheets using the wet transfer method. The specific steps are as follows: The SiO2/Si substrate with the materials was placed on an adhesive machine, and 0.2 g/mL PMMA solution was dropped onto its surface. The adhesive parameters were set to 3000 rpm for 60 s. After applying the adhesive, the substrate was placed on a heating table at 110 °C for 3 min to improve adhesion between the materials and the PMMA film. The substrate was then immersed in a 10 M KOH solution to etch away the SiO2 layer, allowing the PMMA film with the materials to detach and float off spontaneously. After removing the residual KOH solution, the PMMA film was washed with deionized water. The clean copper sheet and PMMA film were placed on a heating table at 80 °C to ensure the materials adhered completely to the copper sheet. Finally, the PMMA glue was removed by placing the film in acetone, leaving only the materials adhered to the copper sheet.

3.3. Material Characterization

The morphology and thickness of the nanosheets were observed and measured using atomic force microscopy (AFM, MFP-3D, Oxford Instruments, Santa Barbara, CA, USA). Zone-selective electron diffraction (SAED) tests were conducted with transmission electron microscopy (TEM, JEOL-2100Plus, Japan Electronics, Tokyo, Japan). Raman spectroscopy and second harmonic generation (SHG) tests were carried out using a high-resolution Raman spectrometer (HR Evolution, Horiba Jobin Yvon, Glasgow, UK). Elemental composition was analyzed using X-ray photoelectron spectroscopy (XPS, ESCALAB 250Xi, Thermo Fisher Scientific, Waltham, MA, USA).

3.4. Electrochemical Tests

Current–voltage (I–V) tests were performed using a semiconductor parameter analyzer (B1500A, KEYSIGHT, Santa Rosa, CA, USA). Catalytic hydrogen evolution tests were conducted using an electrochemical workstation with a three-electrode system (CHI 760E, Shanghai Chenhua Instrument Co., Ltd., Shanghai, China), which consisted of a reference electrode, counter electrode, and working electrode. A mercury–mercuric oxide (Hg/HgO) electrode was used as the reference electrode, and a graphite rod was used as the counter electrode. Before the tests, 200 mL of 1 M KOH solution was prepared as the electrolyte. The actual contact area between the working electrode and the electrolyte was 0.5 cm2. All tests were conducted at room temperature.

4. Conclusions

In this study, an improved bidirectional gas flow system was utilized to fabricate WS2 nanosheets with high-density edge active sites in a screwed cone shape and single layer on SiO2/Si substrates via the PVD method. Through I-V tests on two types of WS2, it was found that due to the screwed structure, where electron transport occurs through the edge sites of the helix, there is no need to overcome the van der Waals potential barriers between layers. As a result, the current of screwed WS2 is 104 times higher than that of monolayer WS2, significantly enhancing charge transport efficiency. In contrast to conventional hydrogen evolution reaction (HER) testing in acidic media, this study employed an alkaline environment to prevent edge site deactivation caused by excessive H⁺ adsorption. The screwed WS2 structure significantly increases the density of edge sites and sulfur vacancies by introducing dislocations, interlayer distortions, and stacking defects. These sites exhibit unsaturated coordination of sulfur atoms, making them more prone to adsorbing hydrogen atoms, and thus acting as efficient reaction centers. The interlayer coupling and defects in the screwed structure enhance electron conductivity, accelerate charge transfer, and reduce electrochemical polarization. Furthermore, the screwed structure improves mechanical strength, reducing the risk of structural collapse during the catalytic process.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal15050496/s1, Figure S1: Schematic illustration of preparation of WS2 nanosheets. Figure S2: XPS characterization of WS2 nanosheets. Figure S3: Raman characterization of WS2 nanosheets. Figure S4: AFM characterization of monolayer WS2. Figure S5: Nonlinear optical response of monolayer WS2 via second harmonic generation. Figure S6: Electric double-layer capacitance testing of screwed WS2 and monolayer WS2.

Author Contributions

D.H.: investigation, original draft writing, formal analysis, methodology; C.S., Y.W. and F.Z.: project administration, validation, supervision; D.H. and H.L.: writing; Y.L., L.S. and C.L.: supervision, validation, reviewing, editing; W.Z.: SHG testing; C.L. and H.L.: funding resource; W.Z. and H.L.: designing the experiments and analyzing the results. All authors discussed results, commented on them, and edited. All authors have read and agreed to the published version of the manuscript.

Funding

This project was funded by the Natural Science Foundation of Hebei Province (E2024201020), the National Natural Science Foundation of China (51802089), the Science and Technology Project of Hebei Education Department (BJK2024169), and the Advanced Talents Incubation Program of the Hebei University (521100222057).

Data Availability Statement

Data are contained within the article or Supplementary Materials.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Xiong, G.W.; Chen, Y.K.; Zhou, Z.Q.; Liu, F.; Liu, X.Y.; Yang, L.J.; Liu, Q.L.; Sang, Y.H.; Liu, H.; Zhang, X.L.; et al. Rapid synthesis of various electrocatalysts on Ni foam using a universal and facile induction heating method for efficient water splitting. Adv. Funct. Mater. 2021, 31, 2009580. [Google Scholar] [CrossRef]
  2. Zhang, H.L.; Li, J.J.; Chen, Y.; Wu, J.Y.; Wang, K.; Chen, L.J.; Wang, Y.; Jiang, X.W.; Liu, Y.Y.; Wu, Y.L.; et al. Magneto-electrically enhanced intracellular catalysis of FePt-FeC heterostructures for chemodynamic therapy. Adv. Mater. 2021, 33, 2100472. [Google Scholar] [CrossRef] [PubMed]
  3. Kibsgaard, J.; Chorkendorff, I. Considerations for the scaling-up of water splitting catalysts. Nat. Energy 2019, 4, 430–433. [Google Scholar] [CrossRef]
  4. Dotan, H.; Landman, A.; Sheehan, S.W.; Malviya, K.D.; Shter, G.E.; Grave, D.A.; Arzi, Z.; Yehudai, N.; Halabi, M.; Gal, N.; et al. Decoupled hydrogen and oxygen evolution by a two-step electrochemical-chemical cycle for efficient overall water splitting. Nat. Energy 2019, 4, 786–795. [Google Scholar] [CrossRef]
  5. Chang, K.; Mei, Z.W.; Wang, T.; Kang, Q.; Ouyang, S.X.; Ye, J.H. MoS2/graphene cocatalyst for efficient photocatalytic H2 evolution under visible light irradiation. ACS Nano 2014, 8, 7078–7087. [Google Scholar] [CrossRef]
  6. Karunadasa, H.I.; Montalvo, E.; Sun, Y.J.; Majda, M.; Long, J.R.; Chang, C.J. A molecular MoS2 edge site mimic for catalytic hydrogen generation. Science 2012, 335, 698–702. [Google Scholar] [CrossRef] [PubMed]
  7. Laursen, A.B.; Kegnæs, S.; Dahl, S.; Chorkendorff, I. Molybdenum sulfides—Efficient and viable materials for electro-and photoelectrocatalytic hydrogen evolution. Energy. Environ. Sci. 2012, 5, 5577–5591. [Google Scholar] [CrossRef]
  8. Li, Z.H.; Jiang, Z.Z.; Zhou, W.D.; Chen, M.Y.; Su, M.X.; Luo, X.F.; Yu, T.; Yuan, C.L. MoS2 nanoribbons with a prolonged photoresponse lifetime for enhanced visible light photoelectrocatalytic hydrogen evolution. Inorg. Chem. 2021, 60, 1991–1997. [Google Scholar] [CrossRef]
  9. Voiry, D.; Fullon, R.; Yang, J.; de Carvalho Castro e Silva, C.; Kappera, R.; Bozkurt, I.; Kaplan, D.; Lagos, M.J.; Batson, P.E.; Gupta, G.; et al. The role of electronic coupling between substrate and 2D MoS2 nanosheets in electrocatalytic production of hydrogen. Nat. Mater. 2016, 15, 1003–1009. [Google Scholar] [CrossRef]
  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, 100–102. [Google Scholar] [CrossRef]
  11. Bockris, J.O.M.; Parsons, R.; Rosenberg, H. The kinetics of hydrogen evolution. J. Chem. Phys. 1950, 18, 762–763. [Google Scholar] [CrossRef]
  12. Voiry, D.; Mohite, A.; Chhowalla, M. Phase engineering of transition metal dichalcogenides. Chem. Soc. Rev. 2015, 44, 2702–2712. [Google Scholar] [CrossRef] [PubMed]
  13. Zhuo, H.Y.; Zhang, X.; Liang, J.X.; Yu, Q.; Xiao, H.; Li, J. Theoretical understandings of graphene-based metal single-atom catalysts: Stability and catalytic performance. Chem. Rev. 2020, 120, 12315–12341. [Google Scholar] [CrossRef]
  14. Lin, X.Q.; Liu, Y.Y.; Wang, K.; Wei, C.; Zhang, W.; Yan, Y.L.; Li, Y.J.; Yao, J.N.; Zhao, Y.S. Two-Dimensional Pyramid-Like WS2 Layered Structures for Highly Efficient Edge Second-Harmonic Generation. ACS Nano 2018, 12, 689–696. [Google Scholar] [CrossRef] [PubMed]
  15. Wei, C.; Wu, W.Z.; Li, H.; Lin, X.C.; Wu, T.; Zhang, Y.D.; Xu, Q.; Zhang, L.P.; Zhu, Y.H.; Yang, X.; et al. Atomic plane-vacancy engineering of transition-metal dichalcogenides with enhanced hydrogen evolution capability. ACS Appl. Mater. Interfaces 2019, 11, 25264–25270. [Google Scholar] [CrossRef]
  16. Liu, J.; Gao, B.; Qin, W.J.; Jiang, J.Y.; Li, M.X.; Wei, C.; Xu, Q. Selective sulfuration of two-dimensional nonlayered MoO2 nanosheets for high-current-density hydrogen evolution. ACS Appl. Energy Mater. 2022, 5, 10483–10489. [Google Scholar] [CrossRef]
  17. Li, W.F.; Liu, J.H.; Wu, R.R.; Qiao, A.; Li, S.H.; Wei, C. Ultrasensitive and Light Erasable Surface-Enhanced Raman Scattering Substrates Based on Au-MoO2 Heterostructures. Talanta 2025, 287, 127669. [Google Scholar] [CrossRef]
  18. Li, W.F.; Wu, R.R.; Shang, X.Q.; Li, S.H.; Tian, Q.Y.; Xu, Q.; Wei, C. Light erasable surface-enhanced Raman scattering substrates based on the metallic molybdenum dioxide nanospheres. Sens. Actuators B Chem. 2024, 409, 135576. [Google Scholar] [CrossRef]
  19. Garrett, J.E.; Geisert, R.D.; Zavy, M.T.; Morgan, G.L. Evidence for maternal regulation of early conceptus growth and development in beef cattle. Reproduction 1988, 84, 437–446. [Google Scholar] [CrossRef]
  20. Bollinger, M.V.; Lauritsen, J.V.; Jacobsen, K.W.; Nørskov, J.K.; Helveg, S.; Besenbacher, F. One-dimensional metallic edge states in MoS2. Phys. Rev. Lett. 2001, 87, 196803. [Google Scholar] [CrossRef]
  21. Ye, G.; Gong, Y.; Lin, J.; Li, B.; He, Y.M.; Pantelides, S.T.; Zhou, W.; Vajtai, R.; Ajayan, P.M. Defects engineered monolayer MoS2 for improved hydrogen evolution reaction. Nano Lett. 2016, 16, 1097–1103. [Google Scholar] [CrossRef]
  22. Zhou, W.D.; Chen, M.Y.; Guo, M.M.; Hong, A.J.; Yu, T.; Luo, X.F.; Yuan, C.L.; Lei, W.; Wang, S.G. Magnetic enhancement for hydrogen evolution reaction on ferromagnetic MoS2 catalyst. Nano Lett. 2020, 20, 2923–2930. [Google Scholar] [CrossRef] [PubMed]
  23. Tao, B.L.; Unwin, P.R.; Bentley, C.L. Nanoscale variations in the electrocatalytic activity oflayered transition-metal dichalcogenides. J. Phys. Chem. C 2019, 124, 789–798. [Google Scholar] [CrossRef]
  24. Lukowski, M.A.; Daniel, A.S.; English, C.R.; Meng, F.; Forticaux, A.; Hamers, R.J.; Jin, S. Highly active hydrogen evolution catalysis from metallic WS2 nanosheets. Energy Environ. Sci. 2014, 7, 2608–2613. [Google Scholar] [CrossRef]
  25. Tsai, C.; Chan, K.; Nørskov, J.K.; Abild-Pedersen, F. Theoretical insights into the hydrogen evolution activity of layered transition metal dichalcogenides. Surf. Sci. 2015, 640, 133–140. [Google Scholar] [CrossRef]
  26. Kalinin, S.V.; Dyck, O.; Balke, N.; Neumayer, S.; Tsai, W.Y.; Vasudevan, R.; Lingerfelt, D.; Ahmadi, M.; Ziatdinov, M.; McDowell, M.T.; et al. Toward Electrochemical Studies on the Nanometer and Atomic Scales: Progress, Challenges, and Opportunities. ACS Nano 2019, 13, 9735–9780. [Google Scholar] [CrossRef]
  27. Zhou, H.Q.; Yu, F.; Sun, J.Y.; He, R.; Wang, Y.M.; Guo, C.F.; Wang, F.; Lan, Y.C.; Ren, Z.F.; Chen, S. Highly active and durable self-standing WS2/graphene hybrid catalysts for the hydrogen evolution reaction. J. Mater. Chem. A 2016, 4, 9472–9476. [Google Scholar] [CrossRef]
  28. Yu, S.H.; Chen, W.Z.; Wang, H.Y.; Pan, H.; Chua, D.H.C. Highly stable tungsten disulfide supported on a self-standing nickel phosphide foam as a hybrid electrocatalyst for efficient electrolytic hydrogen evolution. Nano Energy 2019, 55, 193–202. [Google Scholar] [CrossRef]
  29. Jiang, Z.Z.; Zhou, W.D.; Hong, A.J.; Guo, M.M.; Yuan, C.L. MoS2 Moiré superlattice for hydrogen evolution reaction. ACS Energy Lett. 2019, 4, 2830–2835. [Google Scholar] [CrossRef]
  30. Xie, L.; Wang, L.; Zhao, W.; Liu, S.; Huang, W.; Zhao, Q. WS2 moiré superlattices derived from mechanical flexibility for hydrogen evolution reaction. Nat. Commun. 2021, 12, 5070. [Google Scholar] [CrossRef]
  31. Liu, L.L.; Sun, Y.H.; Cui, X.Q.; Qi, K.; He, X.; Bao, Q.L.; Ma, W.L.; Lu, J.; Fang, H.Y.; Zhang, P.; et al. Bottom-up growth of homogeneous Moiré superlattices in bismuth oxychloride spiral nanosheets. Nat. Commun. 2019, 10, 4472. [Google Scholar] [CrossRef] [PubMed]
  32. Sarma, P.V.; Kayal, A.; Sharma, C.H.; Thalakulam, M.; Mitra, J.; Shaijumon, M.M. Electrocatalysis on edge-rich spiral WS2 for hydrogen evolution. ACS Nano 2019, 13, 10448–10455. [Google Scholar] [CrossRef]
  33. Fan, X.P.; Jiang, Y.; Zhuang, X.J.; Liu, H.J.; Xu, T.; Zheng, W.H.; Fan, P.; Li, H.L.; Wu, X.P.; Zhu, X.L.; et al. Broken symmetry induced strong nonlinear optical effects in spiral WS2 nanosheets. ACS Nano 2017, 11, 4892–4898. [Google Scholar] [CrossRef]
  34. Barman, P.K.; Sarma, P.V.; Shaijumon, M.M.; Kini, R.N. High degree of circular polarization in WS2 spiral nanostructures induced by broken symmetry. Sci. Rep. 2019, 9, 2784. [Google Scholar] [CrossRef]
  35. Xu, F.B.; Yu, H.; Sadrzadeh, A.; Yakobson, B.I. Riemann surfaces of carbon as graphene nanosolenoids. Nano Lett. 2016, 16, 34–39. [Google Scholar] [CrossRef]
  36. Parzinger, E.; Mitterreiter, E.; Stelzer, M.; Kreupl, F.; Ager, J.W.; Holleitner, A.W.; Wurstbauer, U. Hydrogen evolution activity of individual mono-, bi-, and few-layer MoS2 towards photocatalysis. Appl. Mater. Today 2017, 8, 132–140. [Google Scholar] [CrossRef]
  37. Su, M.; Zhou, W.D.; Liu, L.; Chen, M.Y.; Jiang, Z.Z.; Luo, X.F.; Yang, Y.; Yu, T.; Lei, W.; Yuan, C.L. Micro eddy current facilitated by screwed MoS2 structure for enhanced hydrogen evolution reaction. Adv. Funct. Mater. 2022, 32, 2111067. [Google Scholar] [CrossRef]
  38. Zheng, Y.; Jiao, Y.; Zhu, Y.; Li, L.H.; Han, Y.; Chen, Y.; Du, A.; Jaroniec, M.; Qiao, S.Z. Hydrogen evolution by a metal-free electrocatalyst. Nat. Commun. 2014, 5, 3783. [Google Scholar] [CrossRef] [PubMed]
  39. Pizzagalli, L.; Belabbas, I.; Kioseoglou, J.; Chen, J. First-principles calculations of threading screw dislocations in AlN and InN. Phys. Rev. Mater. 2018, 2, 064607. [Google Scholar] [CrossRef]
  40. Zhan, H.F.; Zhang, G.; Yang, C.H.; Gu, Y.T. Breakdown of Hooke’s law at the nanoscale-2D material-based nanosprings. Nanoscale 2018, 10, 18961–18968. [Google Scholar] [CrossRef]
  41. Sun, H.B.; Kong, X.; Park, H.; Liu, F.N.; Lee, Z.H.; Ding, F. Spiral growth of adlayer graphene. Adv. Mater. 2022, 34, 2107587. [Google Scholar] [CrossRef] [PubMed]
  42. Tay, R.Y.; Park, H.J.; Lin, J.J.; Ng, Z.K.; Jing, L.; Li, H.L.; Zhu, M.M.; Tsang, S.H.; Lee, Z.H.; Teo, E.H.T. Concentric and Spiral Few-Layer Graphene: Growth Driven by Interfacial Nucleation vs Screw Dislocation. Chem. Mater. 2018, 30, 6858. [Google Scholar] [CrossRef]
  43. Kobayashi, Y.; Sasaki, S.; Mori, S.; Hibino, H.; Liu, Z.; Watanabe, K.; Taniguchi, T.; Suenaga, K.; Maniwa, Y.; Miyata, Y. Growth and optical properties of high-quality monolayer WS2 on graphite. ACS Nano 2015, 9, 4056–4063. [Google Scholar] [CrossRef]
  44. Cain, J.D.; Shi, F.; Wu, J.; Dravid, V.P. Growth mechanism of transition metal dichalcogenide monolayers: The role of self-seeding fullerene nuclei. ACS Nano 2016, 10, 5440–5445. [Google Scholar] [CrossRef] [PubMed]
  45. Shearer, M.; Samad, L.; Zhang, Y.; Zhao, Y.Z.; Puretzky, A.; Eliceiri, K.W.; Wright, J.C.; Hamers, R.J.; Jin, S. Complex and noncentrosymmetric stacking of layered metal dichalcogenide materials created by screw dislocations. J. Am. Chem. Soc. 2017, 139, 3496–3504. [Google Scholar] [CrossRef] [PubMed]
  46. Zhao, Y.Z.; Zhang, C.Y.; Kohler, D.D.; Scheeler, J.M.; Wright, J.C.; Voyles, P.M.; Jin, S. Supertwisted spirals of layered materials enabled by growth on non-Euclidean surfaces. Science 2020, 370, 442–445. [Google Scholar] [CrossRef]
  47. Zhang, L.M.; Liu, K.H.; Wong, A.B.; Kim, J.; Hong, X.P.; Liu, C.; Cao, T.; Louie, S.G.; Wang, F.; Yang, P.D. Three-dimensional spirals of atomic layered MoS2. Nano Lett. 2014, 14, 6418–6423. [Google Scholar] [CrossRef]
  48. Zhao, Y.; Jin, S. Stacking and twisting of layered materials enabled by screw dislocations and non-Euclidean surfaces. Acc. Mater. Res. 2022, 3, 369–378. [Google Scholar] [CrossRef]
  49. Han, W.Q.; Liu, Z.H.; Pan, Y.B.; Guo, G.N.; Zou, J.X.; Xia, Y.; Peng, Z.M.; Li, W.; Dong, A.G. Designing champion nanostructures of tungsten dichalcogenides for electrocatalytic hydrogen evolution. Adv. Mater. 2020, 32, 2002584. [Google Scholar] [CrossRef]
  50. Ouyang, Y.L.; Li, W.; Zhang, C.W.; Wang, X.Y.; Ma, L.; Wang, H.; Lin, H.; Zou, G.F.; Zhu, J.T. Hydroxide Modified Synthesis of Atomically-Doped Photoluminescent WS2 Monolayers. Adv. Opt. Mater. 2024, 12, 2302596. [Google Scholar] [CrossRef]
  51. Li, Z.J.; Zhai, L.; Zhang, Q.H.; Zhai, W.; Li, P.; Chen, B.; Chen, C.S.; Yao, Y.; Ge, Y.Y.; Yang, H.; et al. 1T′-transition metal dichalcogenide monolayers stabilized on 4H-Au nanowires for ultrasensitive SERS detection. Nat. Mater. 2024, 23, 1355–1362. [Google Scholar] [CrossRef]
  52. Zhang, S.H.; Liu, H.; Zhang, F.; Zheng, X.M.; Zhang, X.Z.; Zhang, B.H.; Zhang, T.; Ao, Z.K.; Zhang, X.Y.; Liu, H.; et al. Controllable Synthesis of WSe2–WS2 Lateral Heterostructures via Atomic Substitution. ACS. Nano. 2024, 18, 30321–30331. [Google Scholar] [CrossRef] [PubMed]
  53. Sebastian, A.; Pendurthi, R.; Choudhury, T.H.; Redwing, J.M.; Das, S. Benchmarking monolayer MoS2 and WS2 field-effect transistors. Nat. Commun. 2021, 12, 693. [Google Scholar] [CrossRef] [PubMed]
  54. Zhao, W.W.; Cai, S.; Wei, X.; Zheng, T.; Xu, X.; Zafar, A.; Liu, H.W.; Yu, T.; Lu, J.P.; Chen, Y.F. The thinnest light disk: Rewritable data storage and encryption on WS2 monolayers. Adv. Funct. Mater. 2021, 31, 2103140. [Google Scholar] [CrossRef]
  55. Zeng, Z.X.S.; Sun, X.X.; Zhang, D.L.; Zheng, W.H.; Fan, X.P.; He, M.; Xu, T.; Sun, L.T.; Wang, X.; Pan, A.L. Controlled vapor growth and nonlinear optical applications of large-area 3R phase WS2 and WSe2 atomic layers. Adv. Funct. Mater. 2019, 29, 1806874. [Google Scholar] [CrossRef]
  56. Aoki, S.; Dong, Y.; Wang, Z.Q.; Huang, X.S.W.; Itahashi, Y.M.; Ogawa, N. Giant Modulation of the Second Harmonic Generation by Magnetoelectricity in Two-Dimensional Multiferroic CuCrP2S6. Adv. Mater. 2024, 36, 2312781. [Google Scholar] [CrossRef]
  57. Guo, S.Q.; Wang, Y.Y.; Zhang, J.Y. Realization of valley polarization in monolayer WS2 via 3d transition metal atom adsorption. J. Phys. D Appl. Phys. 2020, 53, 384001. [Google Scholar] [CrossRef]
  58. Zhang, Q.Y.; Hou, L.L.; Shautsova, V.; Warner, H. Ultrathin all-2D lateral diodes using top and bottom contacted laterally spaced graphene electrodes to WS2 semiconductor monolayers. ACS Appl. Mater. Interfaces. 2023, 15, 18012–18021. [Google Scholar] [CrossRef]
Figure 1. (a) Monolayer WS2 growth simulation. (b) Screwed WS2 growth simulation. (c,d) Optical microscope morphology of monolayer WS2 and screwed WS2 nanosheets. (e) Screwed WS2 AFM topography. (f) Screwed WS2 thickness information. (g) Observation of edge of screwed WS2 by AFM. (h) SAED of screwed WS2.
Figure 1. (a) Monolayer WS2 growth simulation. (b) Screwed WS2 growth simulation. (c,d) Optical microscope morphology of monolayer WS2 and screwed WS2 nanosheets. (e) Screwed WS2 AFM topography. (f) Screwed WS2 thickness information. (g) Observation of edge of screwed WS2 by AFM. (h) SAED of screwed WS2.
Catalysts 15 00496 g001
Figure 2. (a) Spectral profiles of SHG signals from screwed WS2 and monolayer WS2 (illustrations are screwed (Left) and monolayer (Right) WS2SHG mapping). (b) Power-dependent SHG intensity. (c) SHG intensity vs. laser power in logarithmic coordinates extracted from panel (b). (d) Spectral plots of polarized SHG signals of screwed WS2.
Figure 2. (a) Spectral profiles of SHG signals from screwed WS2 and monolayer WS2 (illustrations are screwed (Left) and monolayer (Right) WS2SHG mapping). (b) Power-dependent SHG intensity. (c) SHG intensity vs. laser power in logarithmic coordinates extracted from panel (b). (d) Spectral plots of polarized SHG signals of screwed WS2.
Catalysts 15 00496 g002
Figure 3. (a,b) Electrical properties of screwed WS2 and monolayer WS2. (c) Volt-ampere characteristic curve of WS2 (the illustration is a schematic diagram of electrochemical testing simulation). (d) Tafel curve of WS2. (e) Impedance spectroscopy of WS2. (f) Chronoamperometry tests were performed on screwed WS2 at a constant polarization potential of −0.36 V vs. RHE, while monolayer WS2 was evaluated at −0.4 V vs. RHE under identical conditions.
Figure 3. (a,b) Electrical properties of screwed WS2 and monolayer WS2. (c) Volt-ampere characteristic curve of WS2 (the illustration is a schematic diagram of electrochemical testing simulation). (d) Tafel curve of WS2. (e) Impedance spectroscopy of WS2. (f) Chronoamperometry tests were performed on screwed WS2 at a constant polarization potential of −0.36 V vs. RHE, while monolayer WS2 was evaluated at −0.4 V vs. RHE under identical conditions.
Catalysts 15 00496 g003
Figure 4. (a) Schematic diagram of hydrogen evolution mechanism of screwed WS2. (b) Schematic diagram of hydrogen evolution mechanism of monolayer WS2.
Figure 4. (a) Schematic diagram of hydrogen evolution mechanism of screwed WS2. (b) Schematic diagram of hydrogen evolution mechanism of monolayer WS2.
Catalysts 15 00496 g004
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

Hu, D.; Sun, C.; Wang, Y.; Zhao, F.; Li, Y.; Song, L.; Lv, C.; Zheng, W.; Li, H. Growth of Two-Dimensional Edge-Rich Screwed WS2 with High Active Site Density for Accelerated Hydrogen Evolution. Catalysts 2025, 15, 496. https://doi.org/10.3390/catal15050496

AMA Style

Hu D, Sun C, Wang Y, Zhao F, Li Y, Song L, Lv C, Zheng W, Li H. Growth of Two-Dimensional Edge-Rich Screwed WS2 with High Active Site Density for Accelerated Hydrogen Evolution. Catalysts. 2025; 15(5):496. https://doi.org/10.3390/catal15050496

Chicago/Turabian Style

Hu, Dengchao, Chaocheng Sun, Yida Wang, Fade Zhao, Yubao Li, Limei Song, Cuncai Lv, Weihao Zheng, and Honglai Li. 2025. "Growth of Two-Dimensional Edge-Rich Screwed WS2 with High Active Site Density for Accelerated Hydrogen Evolution" Catalysts 15, no. 5: 496. https://doi.org/10.3390/catal15050496

APA Style

Hu, D., Sun, C., Wang, Y., Zhao, F., Li, Y., Song, L., Lv, C., Zheng, W., & Li, H. (2025). Growth of Two-Dimensional Edge-Rich Screwed WS2 with High Active Site Density for Accelerated Hydrogen Evolution. Catalysts, 15(5), 496. https://doi.org/10.3390/catal15050496

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