Next Article in Journal
Recent Development of Tunable Optical Devices Based on Liquid
Next Article in Special Issue
Facile Controlled Synthesis of Pd-ZnO Nanostructures for Nitrite Detection
Previous Article in Journal
Bimetallic Lanthanum-Cerium-Loaded HZSM-5 Composite for Catalytic Deoxygenation of Microalgae-Hydrolyzed Oil into Green Hydrocarbon Fuels
Previous Article in Special Issue
Transition Metal Substituted Barium Hexaferrite-Modified Electrode: Application as Electrochemical Sensor of Acetaminophen
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Morphology-Controlled Synthesis of V1.11S2 for Electrocatalytic Hydrogen Evolution Reaction in Acid Media

1
College of Chemical Engineering, Sichuan University of Science & Engineering, Zigong 643000, China
2
School of Mechanical Engineering, Chengdu University, Chengdu 610106, China
3
School of Chemical Engineering, Sichuan University, Chengdu 610065, China
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(22), 8019; https://doi.org/10.3390/molecules27228019
Submission received: 19 October 2022 / Revised: 15 November 2022 / Accepted: 15 November 2022 / Published: 18 November 2022
(This article belongs to the Special Issue Nanomaterials for Electrocatalytic Applications)

Abstract

:
High-performance low-cost catalysts are in high demand for the hydrogen evolution reaction (HER). In the present study, we reported that V1.11S2 materials with flower-like, flake-like, and porous morphologies were successfully synthesized by hydrothermal synthesis and subsequent calcination. The effects of morphology on hydrogen evolution performance were studied. Results show that flower-like V1.11S2 exhibits the best electrocatalytic activity for HER, achieving both high activity and preferable stability in 0.5 M H2SO4 solution. The main reason can be ascribed to the abundance of catalytically active sites and low charge transfer resistance.

1. Introduction

Nowadays, the development of green and sustainable energy has become an important research topic due to the depletion of fossil fuels, as well as increasing serious environmental issues [1,2]. Therefore, it is urgent to develop renewable and clean alternatives. Electrochemical water splitting is considered to be a clean and sustainable way to produce hydrogen fuel [3]. Moreover, the abundance of protons in acid electrolytes facilitates the acceleration of the hydrogen evolution reaction [4]. However, the acidic electrolyte can cause severe chemical corrosion of electrolyzers, which limits the use of non-platinum group metals or their compounds as catalysts [5]. In particular, the high cost and insufficient reserves of precious metals have greatly restricted their large-scale commercial applications [6,7]. Therefore, a lot of research has been focused on exploring low-cost electrocatalysts [8,9,10,11,12].
Among the various hydrogen evolution reaction (HER) catalysts, transition metal chalcogenides (TMDs) have made tremendous progress due to their high catalytic activity toward the HER, as well as their low-cost [13,14]. MoS2 is one of the most excellent electrocatalytic materials among transition metal sulfides, and the catalytic activity and mechanism of MoS2 for HER have been widely understood [15,16,17,18]. MoS2 with a two-dimensional (2D) layered structure is known to contain both active edge sites and chemically inert basal plane. Lots of work has been conducted to improve the activity by increasing the edge sites of MoS2 and/or exploiting the inert basal plane to create additional active sites [19,20]. Hexagonal 1T-phase VS2 (1T-VS2) as a group TMDs is a promising HER electrocatalyst. The structure of 1T-VS2 is similar to that of MoS2, which is assembled by stacked S-V-S monolayers via weak van der Waals interaction, which also has excellent structural stability. For the first time, Pan demonstrated by density functional theory calculation that the catalytic performance of single-layer VS2 is equivalent to that of Pt at low hydrogen coverage [21]. Zhang and his colleagues further explained the role of intrinsic point defects in HER activity of monolayer VS2 catalyst [22]. After that, Liang et al. developed a facile hydrothermal calcination method to synthesize self-supported VS2 on carbon paper, which shows excellent HER properties [23]. Qu and his colleagues also prepared VS2 with flower-like morphology, obtaining superior HER performance in acid solution [24].
V1.11S2 phase is one of nonstoichiometric 1T-V1 + XS2 (0 < X< 0.17) with V atoms in the interstitial site between adjacent layers (X is the concentration of V atoms) [25,26]. Both theoretical and experimental results indicate the excellent HER activity of self-intercalated V1.11S2, which shows a much faster proton/electron adsorption and hydrogen release process than the VS2 [26]. Despite these advances, there were few reports focused on the morphology-controlled synthesis of V1.11S2, as well as their effect on HER performance. It is well known that electrocatalytic activities are highly reliant on the catalyst morphology, which is given more edge sites and lowly coordinated surface atoms that often determine the catalytic performance [27].
Herein, different morphologies of V1.11S2 were synthesized by a simple hydrothermal synthesis and subsequent calcination (Figure 1). The electrochemical catalytic properties of the resultant V1.11S2 materials were systematically investigated.

2. Results and Discussion

Figure 2 shows the XRD patterns of the obtained V1.11S2 materials. All the diffraction peaks can be assigned to the V1.11S2 (33–1445) phase without discernible impurities. It was found that both V1.11S2-1 and V1.11S2-2 have a well-crystalline phenomenon. It can be seen from Figure 2 and Figure S1 that the XRD diffraction peaks before and after calcination are quite different, which is mainly due to the transformation of VS4 and VS2 to V1.11S2 at high-temperature conditions [28].
Figure 3 shows the FE-SEM images of V1.11S2 materials. Figure 3a,b display that flower-like V1.11S2 is stacked by a large number of V1.11S2 nanoplates in different directions. It is worth noting that the average radius of a single V1.11S2 nanoflower is about 10 μm. Moreover, the formation of flowerlike V1.11S2 probably involves two steps [29]. First, in a weak alkaline environment, the −SH functional group produced by C2H5NS reacts with the precursor of V to form V-S intermediate complexes followed by decomposing to shape VSx (x = 2, 4) nuclei for further growth. Then, VSx nanoplates are transformed at high temperatures into V1.11S2 nanosheets, which are stacked together to form flower-like structures. Irregular flake V1.11S2 prepared by process B is shown in Figure 3c and Figure S2a. It can be seen from Figure S3b that the precursor obtained in process B is closely stacked by nanosheets, which are dispersed and smaller after calcination. When the solvent change to ethanol, a porous structure can be observed (Figure 3d). Compared with the powder before calcination shown in Figure S3c, the morphology changes dramatically, which is mainly due to the slight solubility of NaVO3 in ethanol solution and the calcination procedure [30]. These results suggest the morphology can be easily controlled by altering the hydrothermal solvent and the source of vanadium. Transmission electron microscope (TEM) measurements were performed to analyze the physical structure of the V1.11S2-1. As depicted in Figure 4a, flower-like V1.11S2-1 can be exfoliated into a nanosheet structure under long−time ice bath ultrasound. The high-magnification TEM image shown in Figure 4b exhibits the periodic lattice fringe pattern, and the inter-planar spacing was measured to be 0.163 nm, which agrees with that of the (110) facet of V1.11S2 (Figure 4b). The corresponding selected area electron diffraction (SAED) pattern also confirmed the crystal structure of the V1.11S2 phase (inset in Figure 4b). Moreover, the EDS pattern in Figure S4 also reveals that V and S can be detected.
To detect the surface chemical state and element composition, X-ray photoelectron spectroscopy (XPS) analysis was performed on V1.11S2 with different morphologies. Investigation of the XPS spectrum clearly shows the presence of V and S (Figure S5). The V 2p spectra can be fitted with two sets of doublet peaks (Figure 5a), and the spectrum of V1.11S2(V 2p) shows two additional broad peaks at a lower binding energy of 513.4, 516.3, 520.9, and 523.8 eV, which can be respectively assigned to V2+2p3/2, V4+2p3/2, V2+2p1/2, and V4+2p1/2 [26]. The peak fitting analysis of S 2p (Figure 5b) confirms the presence of S2− with two peaks located at 160.8 and 162 eV that can be assigned to S2p3/2 and S2p1/2 [26,31]. The combined above-mentioned data indicate that the V1.11S2 materials with different morphologies have been successfully prepared.
The electrocatalytic HER activities of V1.11S2 materials were assessed by linear sweep voltammetry (LSV) using a three-electrode system under 0.5 M H2SO4 acidic aqueous condition. From Figure 6a, V1.11S2-1 exhibits the best catalytic performance, achieving a current density of 10 mA cm−2 with an overpotential of 252 mV, which is superior to the previously reported vanadium sulfide acidic HER electrocatalysts, such as VS2 nanodiscs (420 mV) [32], CFP supported V1.11S2 (259.7 mV) [33], non−templated VS2 (378 mV) [34], and Co-N-doped single-crystal V3S4 nanoparticles (268 mV) [35]. As illustrated in Figure 6b, the calculated Tafel slopes of V1.11S2-1, V1.11S2-2, V1.11S2-3, and 5 wt.% Pt/C are 71.7 mV dec−1, 264.1 mV dec−1, 102.5 mV dec−1, and 30.6 mV dec−1, respectively. It is worth noting that the Tafel slope of 5 wt.% Pt/C is as low as 30.6 mV dec−1, which is consistent with previous studies [36,37]. Therefore, V1.11S2-1 shows lower overpotential and Tafel slope, indicating its high HER activities.
The stability of the catalyst plays an important role in practical application. The stability test of V1.11S2-1 was also carried out by chronopotentiometry test. From Figure S6, the potential remains stable at the current density of 10 mA cm−2. For comparison, the potential of 5 wt.% Pt/C drops dramatically with the extension of test time, which is consistent with previous studies [38,39]. After the chronopotentiometry test, the LSV curves of V1.11S2-1 show a negligible recession phenomenon (Figure S7), suggesting that the catalyst maintains a highly stable catalytic performance. In brief, the above electrochemical test results confirm the flower-like V1.11S2 material has superior electrochemical activity and stability for HER.
According to previous studies, the catalytic active H−adsorption site of the V1.11S2 catalyst is S in the outermost layer [24,40]. Generally, the electrocatalytic activity is highly dependent on the catalyst morphology with more active sites. In order to further clarify the origination of excellent HER performance for V1.11S2 materials, both the electrochemical surface area (ECSA) of the samples were tested. The corresponding current in the applied potential window of 0.06–0.16 V vs. the reversible hydrogen electrode (RHE) should be originated from the charging of the double-layer, and the calculated capacitance (Cdl) should be proportional to the ECSA [41]. As shown in Figure 7 and the corresponding cyclic voltammograms in Figure S8, V1.11S2-1 has higher electric double-layer capacitance (3.4 mF cm−2) than V1.11S2-2 (0.45 mF cm−2) and V1.11S2-2 (1.9 mF cm−2). Moreover, the fitting value R−Squares is listed in Table S1, suggesting V1.11S2-1 has a larger surface area with more exposed active sites. This may be one of the reasons for its high HER performance.
EIS measurements were performed to examine the kinetic differences between V1.11S2 in different morphologies during the electrochemical process [42]. As shown in the illustration in Figure 8, the semicircle in the Nyquist plots was fitted by using the Randles equivalent circuit, in which Rs represents the equivalent series resistance, Rct1 represents the charge transfer resistance of the electrode, and CPE represents the constant phase element [43,44]. It is worth noting that the charge transfer resistance (Rct1) is related to the electrocatalytic kinetics, and a lower value corresponds to a faster reaction rate, which can be quantified from the diameter of the semicircle in the low-frequency zone [45]. Table 1 demonstrates the changing trend of the Rct1 value for V1.11S2 nanomaterials with different morphologies, V1.11S2-1 (49.54 Ω) < V1.11S2-3 (60.8 Ω) < V1.11S2-2 (114.3 Ω), indicating that V1.11S2-1 has better conductivity. Overall, we can conclude that the enhanced catalytic HER activity of flower-like V1.11S2 compared to the other two structures can be accountable for both the abundant catalytically active sites and preferable low charge transfer resistance.

3. Experimental Section

3.1. Materials

Ammonium vanadate (NH4VO3), sodium orthovanadate (Na3VO4·12H2O), sodium metavanadate (NaVO3), vanadyl acetylacetonate (C10H14O5V), thioacetamide (CH3CSNH2), cysteine (C3H7NO2S), ammonia (NH3·H2O), deionized water (H2O), anhydrous ethanol (C₂H₆O), N-methyl pyrrolidone (C5H9NO). The above chemicals and reagents were purchased from Chengdu Kelong Co., Ltd., Chengdu, China. All reagents were used directly without further purification. The commercial 5 wt.% Pt/C catalyst was purchased from Macleans. Carbon paper (TGP-H-060) was purchased from yilongsheng Energy Technology Co., Ltd., Suzhou, China.

3.2. Synthesis of V1.11S2 Materials

The schematic for the synthesis of V1.11S2 with different morphology is shown in Figure 1. In process A, 1 mmol NH4VO3 and 10 mmol CH3CSNH2 were first dissolved in a solution containing 38 mL deionized water and 2 mL NH3·H2O, which was stirred for 1 h to form a uniform solution. Afterward, the prepared solution was transferred to a 50 mL Teflon-lined stainless-steel autoclave and maintained at 160 °C for 24 h. After natural cooling to room temperature, the black precipitates were collected by centrifugation, washed several times with deionized water and absolute ethanol, and dried under a vacuum for 6 h. In process B, 1 mmol NaVO3 was used as a vanadium source, and 15 mmol CH3CSNH2 as a sulfur source, dissolved in 25 mL deionized water; other reaction conditions were the same. In process C, 4 mmol NaVO3, and 24 mmol CH3CSNH2 were dissolved in 25 mL C₂H₆O, and the reaction temperature was raised to 180 °C. All final precipitates were calcined at 400 °C for 2 h to obtain V1.11S2. These V1.11S2 materials obtained by different preparation procedures were labeled as V1.11S2-1, V1.11S2-2, and V1.11S2-3, respectively.

3.3. Materials Characterization

The phase constitutes of the obtained samples were characterized by an X-ray diffractometer (XRD, DX-2700B) with Cu Kα radiation. The microstructure of V1.11S2 with different morphology was examined by a field emission scanning electron microscope (FESEM, FEI Insect F50). The TEM images of V1.11S2−1 were obtained by high-resolution transmission electron microscopy (TEM, FEI Talos F200S Super). The surface valence states and elemental compositions were analyzed using X-ray photoelectron spectroscopy (XPS, Thermo Fischer, ESCALAB Xi+).

3.4. Electrochemical Measurements

All electrochemical data were measured by an electrochemical workstation (CHI660E, CH Instrument, Shanghai, China) with a typical three-electrode electrochemical cell in the acidic electrolyte (0.5 M H2SO4). A graphite rod was used as the counter electrode, and a saturated calomel electrode was used as the reference electrode. The working electrode was prepared as follows: 3 mg catalysts and 50 μL Nafion solution (5 wt.%) were dispersed in 500 μL mixed solvent of deionized water−isopropanol (volume ratio of 3:1), then sonication to form a homogeneous ink solution. Ink with a volume of 15 μL was loaded onto the carbon paper (0.25 cm−2) electrode and dried at ambient temperature. A loading density of about 0.343 mg cm−2 was obtained. Linear sweep voltammograms (LSV) were measured from 0.10 to −0.90 V (vs. RHE) at a scan rate of 5 mV s−1. Electrochemical impedance spectroscopy (EIS) was conducted under −238 mV (vs. RHE) over a frequency range from 100 kHz to 0.01 Hz with a 5mV amplitude potential. The cyclic voltammograms (CV) measurements at various scan rates from 20, 40, 60, 80, and 100 mV s−1 were performed in the potential range of 0.06~0.16 V (vs. RHE) for the electrochemical double−layer capacitance (Cdl) estimation. It should be noted that no iR compensation was applied to our testing data. According to E corrected = E measured i × R s , all potentials are corrected by iR, where i is the test current and Rs is the equivalent series resistance, which is determined by the Nyquist plots fitting.

4. Conclusions

In conclusion, V1.11S2 was successfully synthesized by hydrothermal synthesis and subsequent calcination. The morphology can be easily controlled by altering the hydrothermal solvent and the source of vanadium. The electrocatalysis results show that the flower−like V1.11S2 has the best catalytic activity, which can be ascribed to abundant catalytically active sites and preferable low charge transfer resistance. This research provides us with a strong basis for the morphology dependent of V1.11S2 materials towards HER. Further works will be performed to enhance the intrinsic activity of V1.11S2 and/or supported on conductive substrates such as carbon cloth and metal foam.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27228019/s1, Figure S1: XRD patterns of hydrothermal synthesis precursor powder before annealing; Figure S2: FE-SEM images of (a) V1.11S2-2, and (b) V1.11S2-3 materials; Figure S3: FE-SEM images of precursor powder before annealing, (a) V1.11S2-1, (b) V1.11S2−2, and (c) V1.11S2-3, respectively; Figure S4: EDS pattern of V1.11S2−1 materials; Figure S5: XPS survey spectra of V1.11S2-1, V1.11S2-2, and V1.11S2-3 materials; Figure S6: Chronopotentiometry curve of V1.11S2-1 and Pt/C recorded at the current density of 10 mA cm−2 for a total duration of 5 h; Figure S7: Initial (red line) and after 5 h (purple line) polarization curves of V1.11S2-1 material; Figure S8: Voltammograms of (a) V1.11S2-1, (b) V1.11S2-2 and (c) V1.11S2-3 at various scan rates of 20, 40, 60, 80 and 100 mV s−1, respectively; Table S1: the electric double layer (Cdl) capacitance is obtained by fitting CV curve.

Author Contributions

Methodology, X.L. and W.Y.; Software, X.A., J.Z. and L.X.; Formal analysis, Q.K.; Data curation, Q.C. and S.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of Qingquan Kong grant number: 11832007. This research was funded by the Application and Foundation Research Projects of Qingquan Kong grant number: 2022NSFSC1965, 2022JDRC0085. This research was funded by the Wuliangye Group Industry University Research Cooperation Project of Xiaonan Liu grant number: CXY2021ZR002.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author. The data are not publicly available due to restrictions eg privacy or ethical.

Acknowledgments

Many thanks go to Chenghua Sun from the Swinburne University of Technology who provided discussions and good suggestions.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhou, W.; Jia, J.; Lu, J.; Yang, L.; Hou, D.; Li, G.; Chen, S. Recent developments of carbon−based electrocatalysts for hydrogen evolution reaction. Nano Energy 2016, 28, 29–43. [Google Scholar] [CrossRef]
  2. Chen, Z.; Duan, X.; Wei, W.; Wang, S.; Ni, B.J. Recent advances in transition metal−based electrocatalysts for alkaline hydrogen evolution. J. Mater. Chem. A 2019, 7, 14971–15005. [Google Scholar] [CrossRef] [Green Version]
  3. Zhu, J.; Hu, L.; Zhao, P.; Lee, L.Y.S.; Wong, K.Y. Recent Advances in Electrocatalytic Hydrogen Evolution Using Nanoparticles. Chem. Rev. 2020, 120, 851–918. [Google Scholar] [CrossRef]
  4. Hu, K.; Ohto, T.; Nagata, Y.; Wakisaka, M.; Aoki, Y.; Fujita, J.I.; Ito, Y. Catalytic activity of graphene-covered non-noble metals governed by proton penetration in electrochemical hydrogen evolution reaction. Nat. Commun. 2021, 12, 203. [Google Scholar] [CrossRef]
  5. Yang, C.; Zhao, R.; Xiang, H.; Wu, J.; Zhong, W.; Li, W.; Zhang, Q.; Yang, N.; Li, X. Ni-Activated Transition Metal Carbides for Efficient Hydrogen Evolution in Acidic and Alkaline Solutions. Adv. Energy Mater. 2020, 10, 2002260. [Google Scholar] [CrossRef]
  6. Li, Y.; Wei, X.; Chen, L.; Shi, J. Electrocatalytic Hydrogen Production Trilogy. Angew. Chem. Int. Ed. Engl. 2020, 10, 1–21. [Google Scholar]
  7. Du, H.; Kong, R.M.; Guo, X.; Qu, F.; Li, J. Recent progress in transition metal phosphides with enhanced electrocatalysis for hydrogen evolution. Nanoscale 2018, 10, 21617–21624. [Google Scholar] [CrossRef]
  8. Wang, C.; Tian, B.; Wu, M.; Wang, J. Revelation of the Excellent Intrinsic Activity of MoS2|NiS|MoO3 Nanowires for Hydrogen Evolution Reaction in Alkaline Medium. ACS Appl. Mater. Inter. 2017, 9, 7084–7090. [Google Scholar] [CrossRef]
  9. Zheng, Y.; Jiao, Y.; Jaroniec, M.; Qiao, S.Z. Advancing the Electrochemistry of the Hydrogen−Evolution Reaction through Combining Experiment and Theory. Angew. Chem. Int. Ed. 2015, 54, 52–65. [Google Scholar] [CrossRef]
  10. Eftekhari, A. Electrocatalysts for hydrogen evolution reaction. Int. J. Hydrogen Energy 2017, 42, 11053–11077. [Google Scholar] [CrossRef]
  11. Wang, J.; Cui, W.; Liu, Q.; Xing, Z.; Asiri, A.M.; Sun, X. Recent Progress in Cobalt−Based Heterogeneous Catalysts for Electrochemical Water Splitting. Adv. Mater. 2016, 28, 215–230. [Google Scholar] [CrossRef] [PubMed]
  12. Zhu, W.; Zhang, R.; Qu, F.; Asiri, A.M.; Sun, X. Design and Application of Foams for Electrocatalysis. ChemCatChem 2017, 9, 1721–1743. [Google Scholar] [CrossRef]
  13. Hu, J.; Zhang, C.; Meng, X.; Lin, H.; Hu, C.; Long, X.; Yang, S. Hydrogen evolution electrocatalysis with binary−nonmetal transition metal compounds. J. Mater. Chem. A 2017, 5, 5995–6012. [Google Scholar] [CrossRef]
  14. Li, G.; Zhang, D.; Yu, Y.; Huang, S.; Yang, W.; Cao, L. Activating MoS2 for pH-Universal Hydrogen Evolution Catalysis. J. Am. Chem. Soc. 2017, 139, 16194–16200. [Google Scholar] [CrossRef]
  15. Chen, L.X.; Chen, Z.W.; Wang, Y.; Yang, C.C.; Jiang, Q. Design of Dual-Modified MoS2 with Nanoporous Ni and Graphene as Efficient Catalysts for the Hydrogen Evolution Reaction. ACS Catal. 2018, 8, 8107–8114. [Google Scholar] [CrossRef]
  16. Luo, Z.; Ouyang, Y.; Zhang, H.; Xiao, M.; Ge, J.; Jiang, Z.; Wang, J.; Tang, D.; Cao, X.; Liu, C.; et al. Chemically activating MoS2 via spontaneous atomic palladium interfacial doping towards efficient hydrogen evolution. Nat. Commun. 2018, 9, 2120. [Google Scholar] [CrossRef] [Green Version]
  17. Sun, T.; Wang, J.; Chi, X.; Lin, Y.; Chen, Z.; Ling, X.; Qiu, C.; Xu, Y.; Song, L.; Chen, W.; et al. Engineering the Electronic Structure of MoS2 Nanorods by N and Mn Dopants for Ultra-Efficient Hydrogen Production. ACS Catal. 2018, 8, 7585–7592. [Google Scholar] [CrossRef]
  18. Zhang, H.; Yu, L.; Chen, T.; Zhou, W.; Lou, X.W. Surface Modulation of Hierarchical MoS2 Nanosheets by Ni Single Atoms for Enhanced Electrocatalytic Hydrogen Evolution. Adv. Funct. Mater. 2018, 28, 1807086. [Google Scholar] [CrossRef]
  19. Wan, Y.; Zhang, Z.; Xu, X.; Zhang, Z.; Li, P.; Fang, X.; Zhang, K.; Yuan, K.; Liu, K.; Ran, G.; et al. Engineering active edge sites of fractal-shaped single-layer MoS2 catalysts for high-efficiency hydrogen evolution. Nano Energy 2018, 51, 786–792. [Google Scholar] [CrossRef]
  20. 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] [Green Version]
  21. Pan, H. Metal Dichalcogenides Monolayers: Novel Catalysts for Electrochemical Hydrogen Production. Sci. Rep. 2014, 4, 5348. [Google Scholar] [CrossRef]
  22. Zhang, Y.; Chen, X.; Huang, Y.; Zhang, C.; Li, F.; Shu, H. The Role of Intrinsic Defects in Electrocatalytic Activity of Monolayer VS2 Basal Planes for the Hydrogen Evolution Reaction. J. Phys. Chem. C 2017, 121, 1530–1536. [Google Scholar] [CrossRef]
  23. Liang, H.; Shi, H.; Zhang, D.; Ming, F.; Wang, R.; Zhuo, J.; Wang, Z. Solution Growth of Vertical VS2 Nanoplate Arrays for Electrocatalytic Hydrogen Evolution. Chem. Mater. 2016, 28, 5587–5591. [Google Scholar] [CrossRef]
  24. Qu, Y.; Shao, M.; Shao, Y.; Yang, M.; Xu, J.; Kwok, C.T.; Shi, X.; Lu, Z.; Pan, H. Ultra-high electrocatalytic activity of VS2 nanoflowers for efficient hydrogen evolution reaction. J. Mater. Chem. A 2017, 5, 15080–15086. [Google Scholar] [CrossRef]
  25. Moutaabbid, H.; Le Godec, Y.; Taverna, D.; Baptiste, B.; Klein, Y.; Loupias, G.; Gauzzi, A. High-Pressure Control of Vanadium Self−Intercalation and Enhanced Metallic Properties in 1T-V1+xS2 Single Crystals. Inorg. Chem. 2016, 55, 6481–6486. [Google Scholar] [CrossRef]
  26. Yang, M.; Cao, L.; Wang, Z.; Qu, Y.; Shang, C.; Guo, H.; Xiong, W.; Zhang, J.; Shi, R.; Zou, J.; et al. Vanadium self-intercalated C/V1.11S2 nanosheets with abundant active sites for enhanced electro-catalytic hydrogen evolution. Electrochimi. Acta 2019, 300, 208–216. [Google Scholar] [CrossRef]
  27. Yang, H.G.; Sun, C.H.; Qiao, S.Z.; Zou, J.; Liu, G.; Smith, S.C.; Cheng, H.M.; Lu, G.Q. Anatase TiO2 single crystals with a large percentage of reactive facets. Nature 2008, 453, 638. [Google Scholar] [CrossRef] [Green Version]
  28. Mohan, P.; Yang, J.; Jena, A.; Shin, H.S. VS2/rGO hybrid nanosheets prepared by annealing of VS4/rGO. J. Solid State Chem. 2015, 224, 82–87. [Google Scholar] [CrossRef]
  29. Chen, X.; Yu, K.; Shen, Y.; Feng, Y.; Zhu, Z. Synergistic effect of MoS2 nanosheets and VS2 for the hydrogen evolution reaction with enhanced humidity−sensing performance. ACS Appl. Mater. Interfaces 2017, 9, 42139–42148. [Google Scholar] [CrossRef] [PubMed]
  30. Li, W.; Huang, J.; Feng, L.; Cao, L.; Liu, Y.; Pan, L. Nano-grain dependent 3D hierarchical VS2 microrods with enhanced intercalation kinetic for sodium storage properties. J. Power Sources 2018, 398, 91–98. [Google Scholar] [CrossRef]
  31. Zhang, J.; Zhang, C.; Wang, Z.; Zhu, J.; Wen, Z.; Zhao, X.; Zhang, X.; Xu, J.; Lu, Z. Synergistic Interlayer and Defect Engineering in VS2 Nanosheets toward Efficient Electrocatalytic Hydrogen Evolution Reaction. Small 2018, 14, 1703098. [Google Scholar] [CrossRef] [PubMed]
  32. Kumar, G.M.; Ilanchezhiyan, P.; Cho, H.D.; Lee, D.J.; Kim, D.Y.; Kang, T.W. Ultrathin VS2 nanodiscs for highly stable electro catalytic hydrogen evolution reaction. Int. J. Energ Res 2020, 44, 811–820. [Google Scholar] [CrossRef]
  33. Yu, S.H.; Tang, Z.; Shao, Y.; Dai, H.; Wang, H.Y.; Yan, J.; Pan, H.; Chua, D.H. In Situ Hybridizing MoS2 Microflowers on VS2 Microflakes in a One−Pot CVD Process for Electrolytic Hydrogen Evolution Reaction. ACS Appl. Energy Mater. 2019, 2, 5799–5808. [Google Scholar] [CrossRef]
  34. Guo, T.; Song, Y.; Sun, Z.; Wu, Y.; Xia, Y.; Li, Y.; Sun, J.; Jiang, K.; Dou, S.; Sun, J. Bio-templated formation of defect-abundant VS2 as a bifunctional material toward high-performance hydrogen evolution reactions and lithium-sulfur batteries. J. Energy Chem. 2020, 42, 34–42. [Google Scholar] [CrossRef] [Green Version]
  35. Cao, L.; Wang, L.; Feng, L.; Kim, J.H.; Du, Y.; Yang Kou, L.D.; Huang, J. Co-N-doped single-crystal V3S4 nanoparticles as pH-universal electrocatalysts for enhanced hydrogen evolution reaction. Electrochim. Acta 2020, 335, 135696. [Google Scholar] [CrossRef]
  36. Shi, Y.; Zhou, Y.; Yang, D.R.; Xu, W.X.; Wang, C.; Wang, F.B.; Xu, J.J.; Xia, X.H.; Chen, H.Y. Energy Level Engineering of MoS2 by Transition-Metal Doping for Accelerating Hydrogen Evolution Reaction. J. Am. Chem. Soc. 2017, 139, 15479–15485. [Google Scholar] [CrossRef]
  37. Zhang, J.; Liu, Y.; Sun, C.; Xi, P.; Peng, S.; Gao, D.; Xue, D. Accelerated Hydrogen Evolution Reaction in CoS2 by Transition-Metal Doping. ACS Energy Letter. 2018, 3, 779–786. [Google Scholar] [CrossRef]
  38. Ye, S.; Luo, F.; Zhang, Q.; Zhang, P.; Xu, T.; Wang, Q.; He, D.; Guo, L.; Zhang, Y.; He, C.; et al. Highly stable single Pt atomic sites anchored on aniline-stacked graphene for hydrogen evolution reaction. Energy Environ. Sci. 2019, 12, 1000–1007. [Google Scholar] [CrossRef]
  39. Yu, J.; Guo, Y.; She, S.; Miao, S.; Ni, M.; Zhou Liu, M.W.; Shao, Z. Bigger is Surprisingly Better: Agglomerates of Larger RuP Nanoparticles Outperform Benchmark Pt Nanocatalysts for the Hydrogen Evolution Reaction. Adv. Mater. 2018, 30, 1800047–1800056. [Google Scholar] [CrossRef]
  40. Qu, Y.; Pan, H.; Kwok, C.T.; Wang, Z. A first-principles study on the hydrogen evolution reaction of VS2 nanoribbons. Phys. Chem. Chem. Phys. 2015, 17, 24820–24825. [Google Scholar] [CrossRef]
  41. Liu, Y.; Liang, X.; Gu, L.; Zhang, Y.; Li, G.D.; Zou, X.; Chen, J.S. Corrosion engineering towards efficient oxygen evolution electrodes with stable catalytic activity for over 6000 hours. Nat. Commun. 2018, 9, 2609. [Google Scholar] [CrossRef] [PubMed]
  42. Parishani, M.; Malekfar, R.; Bayat, A.; Gharibi, H. Hydrogen evolution reaction on VS2-NiS2 hybrid nanostructured electrocatalyst in acidic media: A binder-free electrode. J. Iran. Chem. Soc. 2022, 19, 4299–4307. [Google Scholar] [CrossRef]
  43. Feng, T.; Ouyang, C.; Zhan, Z.; Lei, T.; Yin, P. Cobalt doping VS2 on nickel foam as a high efficient electrocatalyst for hydrogen evolution reaction. Int. J. Hydrogen Energy 2022, 47, 10646–10653. [Google Scholar] [CrossRef]
  44. Singh, V.K.; Nakate, U.T.; Bhuyan, P.; Chen, J.; Tran, D.T.; Park, S. Mo/Co doped 1T-VS2 nanostructures as a superior bifunctional electrocatalyst for overall water splitting in alkaline media. J. Mater. Chem. A 2022, 10, 9067–9079. [Google Scholar] [CrossRef]
  45. Lin, Z.; Lin, B.; Wang, Z.; Chen, S.; Wang, C.; Dong, M.; Gao, Q.; Shao, Q.; Ding, T.; Liu, H.; et al. Facile Preparation of 1T/2H Mo(S1-xSex)2 Nanoparticles for Boosting Hydrogen Evolution Reaction. ChemCatChem 2019, 11, 2217–2222. [Google Scholar] [CrossRef]
Figure 1. Schematic for the synthesis of V1.11S2 with different morphology.
Figure 1. Schematic for the synthesis of V1.11S2 with different morphology.
Molecules 27 08019 g001
Figure 2. The XRD patterns of the as-annealed V1.11S2 materials.
Figure 2. The XRD patterns of the as-annealed V1.11S2 materials.
Molecules 27 08019 g002
Figure 3. FE-SEM images of V1.11S2 materials. (a,b) V1.11S2-1, (c) V1.11S2-2, (d) V1.11S2-3.
Figure 3. FE-SEM images of V1.11S2 materials. (a,b) V1.11S2-1, (c) V1.11S2-2, (d) V1.11S2-3.
Molecules 27 08019 g003
Figure 4. (a) TEM images, and (b) high-magnification TEM images of the V1.11S2-1 materials. The inset in Figure (b) is the corresponding SAED pattern.
Figure 4. (a) TEM images, and (b) high-magnification TEM images of the V1.11S2-1 materials. The inset in Figure (b) is the corresponding SAED pattern.
Molecules 27 08019 g004
Figure 5. XPS high-resolution spectra of (a) V 2p and (b) S 2p for the V1.11S2-1, V1.11S2-2, and V1.11S2-3 materials.
Figure 5. XPS high-resolution spectra of (a) V 2p and (b) S 2p for the V1.11S2-1, V1.11S2-2, and V1.11S2-3 materials.
Molecules 27 08019 g005
Figure 6. Electrochemical property of V1.11S2 materials for HER in 0.5 M H2SO4. (a) The iR-corrected polarization curves; (b) Tafel plots of V1.11S2-1, V1.11S2-2, V1.11S2-3, and 5 wt.% Pt/C.
Figure 6. Electrochemical property of V1.11S2 materials for HER in 0.5 M H2SO4. (a) The iR-corrected polarization curves; (b) Tafel plots of V1.11S2-1, V1.11S2-2, V1.11S2-3, and 5 wt.% Pt/C.
Molecules 27 08019 g006
Figure 7. Estimation of Cdl by liner fitting the differences in current density variation (Ja-Jc) at 0.15 V (vs. SCE) as a function of scan rate.
Figure 7. Estimation of Cdl by liner fitting the differences in current density variation (Ja-Jc) at 0.15 V (vs. SCE) as a function of scan rate.
Molecules 27 08019 g007
Figure 8. Nyquist plots for V1.11S2 materials without iR compensation.
Figure 8. Nyquist plots for V1.11S2 materials without iR compensation.
Molecules 27 08019 g008
Table 1. The electrochemical equivalent circuit (EEC) parameters of V1.11S2 nanomaterials with different morphologies were obtained by fitting the Nyquist diagram.
Table 1. The electrochemical equivalent circuit (EEC) parameters of V1.11S2 nanomaterials with different morphologies were obtained by fitting the Nyquist diagram.
CatalystV1.11S2-1V1.11S2-2V1.11S2-3
Rs(Ω)0.36 ± 0.0230.34 ± 0.0140.49 ± 0.021
Rct1 (Ω)49.54 ± 0.48114.3 ± 3.2760.8 ± 1.08
CPE2-T(F)2.03 × 10−3 ± 7.77 × 10−61.84 × 10−4 ± 1.32 × 10−52.15 × 10−3 ± 4.84 × 10−5
CPE2-P(F)0.82 ± 5.33 × 10−30.90 ± 0.0120.69 ± 3.89 × 10−3
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chen, Q.; Tian, S.; Liu, X.; An, X.; Zhang, J.; Xu, L.; Yao, W.; Kong, Q. Morphology-Controlled Synthesis of V1.11S2 for Electrocatalytic Hydrogen Evolution Reaction in Acid Media. Molecules 2022, 27, 8019. https://doi.org/10.3390/molecules27228019

AMA Style

Chen Q, Tian S, Liu X, An X, Zhang J, Xu L, Yao W, Kong Q. Morphology-Controlled Synthesis of V1.11S2 for Electrocatalytic Hydrogen Evolution Reaction in Acid Media. Molecules. 2022; 27(22):8019. https://doi.org/10.3390/molecules27228019

Chicago/Turabian Style

Chen, Qiuyue, Siqi Tian, Xiaonan Liu, Xuguang An, Jingxian Zhang, Longhan Xu, Weitang Yao, and Qingquan Kong. 2022. "Morphology-Controlled Synthesis of V1.11S2 for Electrocatalytic Hydrogen Evolution Reaction in Acid Media" Molecules 27, no. 22: 8019. https://doi.org/10.3390/molecules27228019

Article Metrics

Back to TopTop