Next Article in Journal
Solvent Effect in Catalytic Lignin Hydrogenolysis
Next Article in Special Issue
The High Electrocatalytic Performance of NiFeSe/CFP for Hydrogen Evolution Reaction Derived from a Prussian Blue Analogue
Previous Article in Journal
Hg0 Removal by a Palygorskite and Fly Ash Supported MnO2-CeO2 Catalyst at Low Temperature
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Sulfuration Temperature-Dependent Hydrogen Evolution Performance of CoS2 Nanowires

1
School of Materials Science and Engineering, Chang’an University, Xi’an 710049, China
2
Engineering Research Center of Pavement Materials, Ministry of Education of P.R. China, Chang’an University, Xi’an 710061, China
3
State Key Laboratory for Mechanical Behavior of Materials, Xi’an Jiaotong University, Xi’an 710049, China
*
Authors to whom correspondence should be addressed.
Catalysts 2022, 12(6), 663; https://doi.org/10.3390/catal12060663
Submission received: 20 May 2022 / Revised: 14 June 2022 / Accepted: 14 June 2022 / Published: 17 June 2022

Abstract

:
Densely aligned CoS2 nanowires (NWs) on chemically durable stainless steel fibers felt (SSF) substates were synthesized by thermal sulfuring Co3O4 NWs, which were oxidized from hydrothermal synthesized Co(OH)y(CO3)(1−0.5y)·nH2O NWs precursors. The effect of sulfuration temperature on the composition, morphology, and HER performance of the products was studied in detail. The results show that the high purity together with the enlarged density of active sites given by the twisted morphology of the CoS2 NWs sulfured at 500 °C guarantee its superior hydrogen evolution reaction (HER) performance compared with other samples sulfured at lower temperatures.

1. Introduction

Hydrogen (H2), an environmentally friendly energy carrier with high energy density, has been recognized as a promising alternative to traditional fossil fuels to resolve the global warming and energy crisis issues [1,2,3]. Electrocatalytic water splitting is considered to be a sustainable method for the large-scale production of H2 [4]. To date, the most effective HER catalyst is based on precious metals, especially platinum, which has a very small Tafel slope and high exchange current density [5,6,7]. However, its high price and low nature storage limit its large-scale application, prompting people to develop other effective, abundant, and excellent stability alternative materials [8,9,10].
In recent years, pyrite-phase transition metal dichalcogenides (MX2, where typically M = Fe, Co, Mo or Ni, and X = S or Se) have been extensively investigated for HER applications because of their low cost, simple production, and good electrochemical stability [11,12,13,14]. Among these MX2 materials, cobalt disulfide (CoS2), a typical Co-based material, has attracted great attention due to its metallic conductivity (6.7 × 103 S cm−1 at 300 K) [15], abundant d orbital electrons, and low H adsorption energy, which affirm it as a promising alternative candidate to replace the costly Ptbased electrocatalysts [16]. In the past few years, CoS2 nanostructures with diverse morphology, such as nanosheets [17], nanowires [18], nanoneedles [19], nano polyhedrons [20], and nano flowers [21], have been synthesized via hydrothermal process or the thermal sulfuration of Co-based compounds for HER applications. Those works synthetically demonstrated the promising perspective of CoS2 for future HER application. For the synthesis of CoS2 nanostructures, the direct hydrothermal process usually involves problems of impurity incorporation and hybrid products, which hinder the deployment of the inherent superior physical properties of CoS2 [22]. Thermal sulfuration is a versatility approach to synthesize metal sulfides or sulfur-doped compounds with tunable doping concentration by adjusting reaction temperatures or reaction duration [23,24]. The resulting metal sulfides synthesized by thermal sulfuration usually have relatively high purity because of the controlled vapor environment, and the purity can also be further optimized by adjusting the reaction process conditions. Several studies have shown the advantage of thermal sulfuration to synthesize CoS2 nanostructures with enhanced HER properties by sulfuring Co-based hydroxides [25], cobaltates [26], or polyoxometalates [27].
In this work, densely aligned CoS2 NWs on chemically durable SSF substates were synthesized by thermal sulfuring Co3O4 NWs. The detailed results show that the composition and morphology of the products strongly depend on the sulfuration temperatures. CoS2 NWs with high purity can be obtained at a sulfuration temperature of 500 °C. The high purity, together with the large density of active sites given by its twisted morphology, guarantee superior HER performance compared with other samples ssulfured at lower temperatures.

2. Results and Discussion

As shown in Figure 1a, the SSF substrate used in this work has a large porosity with an average pore size in micron order. The bare SSF fibers have a smooth surface with a uniform diameter of about 10 μm. After the hydrothermal reaction, high-density well aligned Co(OH)y(CO3)(1−0.5y)·nH2O NWs are successfully synthesized on the surface of the SSF substrate with lengths in the range of about 10–15 μm and diameters in the range of 150–200 nm (Figure 1b). After thermal treatment in air, small morphology changes can be found (Figure 1c). Figure 1d–l shows the SEM images of the samples after sulfuration at different temperatures. As is indicated, samples that sulfureted at 300 °C and 400 °C retain the well aligned NW-like morphology (Figure 1d–i), while the sample sulfureted at 500 °C (Figure 1j–l) shows significant morphology changes. The NWs are twisted and intertwined together, forming a reticular structure, leading to an enlarged specific surface area providing more active sites for HER reactions. This obvious morphology change should be related to the accelerated reaction kinetics at high sulfuration temperatures, which results in an obvious anisotropic crystal growth.
Figure 2a shows the XRD patterns of the samples. Before thermal treatment, the sample only show peaks at 43.5°, 50.7°, and 74.7° attributed to SSF substrate (JCPDS:31-0619). No diffraction peaks from Co(OH)y(CO3)(1−0.5y)·nH2O precursors can be found owing to its low crystallinity degree. The thermal oxidized sample shows diffraction peaks at 31.3°, 36.9°, 38.5°, 44.8°, 55.7°, 59.4°, and 65.2° corresponding to the (220), (311), (222), (400), (422), (511), and (440) crystallographic planes of Co3O4 (JCPDS:42-1467) [28]. Apart from those peaks, no other impurity peaks can be indicated, confirming the complete oxidation of Co(OH)y(CO3)(1−0.5y)·nH2O precursors after the thermal oxidation process. The sulfured samples show obvious temperature-dependent phase compositions. As can be identified, the diffraction intensity from CoS2 increases with the increase in sulfuration temperature accompanied by the decreased diffraction intensity from Co3O4. After sulfuration at 500 °C, peaks at 27.9°, 32.3°, 36.2°, 39.8°, 46.3°, 54.9°, and 62.7° are found, attributed to the (111), (200), (210), (211), (220), (311), and (321) crystallographic planes of cubic phase CoS2 (JCPDS 41-1471), respectively [29,30]. The absence of peaks from Co3O4 indicates the full conversion of Co3O4 to CoS2. XPS analysis was used to further investigate the chemical bonding states of the sample sulfureted at 500 °C. As shown in Figure 2b, the survey spectrum shows peaks from S 2p and Co 2p, indicating the formation of CoS2. The C 1s and O 1s peaks are mainly attributed to the surface-adsorbed CO2 and surface oxidations. The Fe 2p peak originated from the Fe elements in the substrate. The high-resolution Co 2p XPS spectrum shows two peaks at binding energies of 778.3 and 793.3 eV corresponding to the spin orbitals of Co 2p3/2 and Co 2p3/2 from Co-S bond, respectively (Figure 2c) [31]. The satellite peaks at 781.0 and 793.4 eV correspond to Co-O bonds owing to surface oxidations. The peaks at 784.9 and 797.3 eV correspond to the vibronic excitation satellite peaks [32]. The high-resolution S 2p spectrum show peaks at 163.5 eV, 162.7 eV, and 168.9 eV, corresponding to Sn2−, S22−, and surface adsorbed sulfates, respectively [18,33].
Figure 3 shows the TEM images of the synthesized CoS2 NWs sulfureted at 500 °C. Consistent with the SEM observations, the NWs show a polycrystalline structure with rough surfaces (Figure 3a). The HRTEM image confirms the well-crystallized CoS2 crystallographic feature, where the lattice spacing is measured to about 0.25 nm, well matched with the (210) crystal planes of CoS2 (Figure 3b). The selected area diffraction pattern confirms the polycrystalline structural feature (Figure 3c). The EDX elemental mapping of CoS2 NWs clearly demonstrates the uniform distribution of Co and S, corroborating the formation of CoS2. A small amount of O element is found to distribute on the surface of CoS2 NWs because of the surface oxidation in the air.
The formation mechanism of the CoS2 NWs can be understood by the chemical reaction shown below that takes place during the synthesis process. In the first step (Equations (1)–(5)), i.e., the hydrothermal synthesis of Co(OH)y(CO3)(1−0.5y)·nH2O precursors, F from NH4F reacts with Co2+ to form [CoF](x−2)− ions, and subsequently reacts with CO32−, OH, and H2O, resulting in the Co(OH)y(CO3)(1−0.5y)·nH2O precursors. According to previous reports, F- is believed to play a key role that determines the NWs morphology. Firstly, the surface-adsorbed F- can reduce the surface activity of the attached crystal grains, facilitating the growth in the direction perpendicular to the substrate [34]. Secondly, F- can also act as a surface reaction catalyst for the growth of Co(OH)y(CO3)(1−0.5y)·nH2O NWs and thereby maintain its stable growth (Equation (5)) [35]. In the second step, the Co(OH)y(CO3)(1−0.5y)·nH2O NW precursors react with O2 in air at a high temperature to form Co3O4 NWs (Equation (6)) [36]. Finally, the Co3O4 NWs can be sulfured to CoS2 in the tube furnace (Equation (7)). At higher sulfuration temperatures, more S gas is generated, and thereby the sulfuration reaction will be accelerated. The grain size of the resulting CoS2 will be enlarged, driven by the accelerated reaction speed and diffusion rate at higher temperatures. The fast anisotropic growth of the CoS2 grains would therefore be the reason for the bent NW morphology.
C o 2 + + x F [ C o F ] ( x 2 )
H 2 N C O N H 2 S + H 2 O 2 N H 3 + C O 2
C O 2 + H 2 O C O 3 2 + 2 H +
N H 3 H 2 O N H 4 + + O H
[ C o F ] ( x 2 ) + ( 1 0.5 y ) C O 3 2 + y O H + n H 2 O C o ( O H ) y ( C O 3 ) ( 1 0.5 y ) n H 2 O + x F
C o ( O H ) y ( C O 3 ) ( 1 0.5 y ) n H 2 O + O 2 2 C o 3 O 4 + 3 ( 2 n + y ) H 2 O + ( 6 3 y ) C O 2
C o 3 O 4 + 6 S 3 C o S 2 + 2 O 2
Figure 4 shows the detailed electrochemical HER performances, which are carried out in 1.0 M KOH solution with a three-electrode system. The IR-corrected LSV curves and the corresponding Tafel slopes are shown in Figure 4a,b, respectively. The commercial Pt/C sample shows an overpotential of 0.106 V to drive an 100 mA·cm−2 (η100) and a Tafel slope of 30.28 mV·dec-1, which are consistent with previous reports [37,38]. The η100 for CoS2 NWs sulfured at 300, 400, and 500 °C are 0.294, 0.279, and 0.224 V, respectively. The η10 for CoS2 NWs sulfured at 300, 400, 500 ℃, and Pt/C are 0.167, 0.152, 0.121, and 0.042 V, respectively. The sample sulfured at 500 °C exhibits the lowest overall potential, indicating its superior HER activity. The sample also shows the lowest Tafel slope (58.15 mV·dec−1) among all the sulfured samples, indicating the fastest HER kinetics. Electrochemical impedance spectroscopy (EIS) measurements were conducted to evaluate the charge transfer kinetics at the interface of the cathode surface and the electrolyte. As shown in Figure 4c, the Nyquist impedance spectra of the samples show approximately equal values to Rs and Rct, indicating the minimal effect of sulfuration temperature on the charge transfer properties of the samples. The electrochemically active surface areas (ECSA) of the sulfured samples were estimated by electrochemical double-layer capacity (Cdl). As shown in Figure 4d, the fitted Cdl values for samples sulfured at 300, 400, and 500 °C are 9.12, 14.34, and 14.90 mF·cm−2, respectively, indicating the largest density of the active sites of the 500 °C sulfured sample.
The electrocatalytic stability of the sample sulfured at 500 °C was studied by the chronoamperometry and continuous cyclic voltammograms methods. Figure 5a shows a plot of the constant voltage and current density versus time curve. A stable current density is maintained during the potentiostatic measurement for more than 10 h. The polarization curves also show little change after 1000 CV cycles (Figure 5b). Moreover, the sample demonstrates minimal morphology change after the long-term chronoamperometry stability tests. Those results solidify the excellent stability of the 500 °C sulfured CoS2/SSF sample.

3. Experimental

3.1. Materials

Cobalt chloride (CoCl2·6H2O) was purchased from Tianjin Damao Chemical Reagent Factory (Tianjin, China), urea (CH4N2O) was purchased from Tianjin Hongyan Chemical Reagent Factory (Tianjin, China), and ammonium fluoride (NH4F) was purchased from Guangdong Guanghua Technology Co., Ltd. (Guangdong, China). Deionized water was used as the solvent. All used reagents were of analytical grade and without further purification. 316L SSF felt was purchased from Dingrun Siwang Manufacturing Co., Ltd. (Hengshui, China).

3.2. Synthesis of the Catalysts

The synthesis procedure for the CoS2 NWs is shown in Figure 6. Firstly, SSF felts with sizes of 1 × 3 cm was ultrasonically cleaned by acetone, ethanol, and water for 15 min each, respectively, and then the washed SSF felts were dried in air at 60 °C for 2 h. Secondly, 1.785 g cobalt chloride, 0.601 g urea, and 0.148 g ammonium fluoride were mixed with 40 mL of deionized water and stirred for 10 min to obtain a uniform solution. Then, the solution with the cleaned SSF substrates was transferred into a Teflon-lined autoclave and heated in an oven at 180 °C for 3 h. After the reaction, the autoclave was cooled down to room temperature naturally. The obtained products were washed with acetone, ethanol, and deionized water, respectively, and dried at 60 °C for 30 min, and Co(OH)y(CO3)(1−0.5y)·nH2O NWs on SSF substrate precursors were obtained.
The obtained CoS2 NWs precursor samples were thermal treated in a furnace at 500 °C for 30 min to obtain the Co3O4/SSF samples. Then, the Co3O4/SSF samples were sulfured in a tube furnace. Sulfur powder (1 g) and the Co3O4/SSF samples were put at the upper and lower vent of the tube furnace, respectively. Argon gas was injected into the tube furnace for protection with a gas flow of 10 sccm. The tube furnace was heated to 300 °C, 400 °C, and 500 °C, with a heating rate of 10 °C/min, respectively. After being sulfured for 1 h, the resulting products were obtained.

3.3. Characterizations

A Bruker D8 advanced powder X-ray diffractometer (XRD) with Cu Kα radiation was used to identify the crystal structure of the prepared samples (Bruker Group, Karlsruhe, Germany). The morphology of the samples was characterized by S-4800 Hitachi field emission scanning electron microscopy (FESEM) (Tokyo, Japan). The microstructure of the samples was characterized by a JEOL JEM-2100Plus transmission electron microscope (TEM) operating at 200 kV (Beijing, Japan). XPS spectra were measured on a Thermo Scientific K-alpha XPS spectrometer, and the binding energies were corrected by referencing the peak of C1s at 284.60 eV (Thermo Fisher Scientific, Waltham, MA, USA). The contact angle measurement of samples using contact angle gauges (JC2000D1, TEOL, Zhongchen Digital Technology Equipment Co., Ltd., Shanghai, China).

3.4. Electrochemical Measurements

Electrochemical measurements were performed with VersaSTAT 3F electrochemical workstation in 1 M KOH aqueous solution (Shaanxi Foreign Trade Import & Export Co., Ltd., Xi’an, China). All electrochemical properties were tested with a typical three-electrode cell at room temperature. The sample, graphite rod, and saturated calomel electrode were used as working, counter, and reference electrodes, respectively. Before the test, the prepared samples were sealed with 704 silica gel to expose a 1 cm2 surface and dried for 12 h. The Pt/C (20%) loaded SSF was also tested for comparison. All the potentials (vs. Hg/HgO) were referenced to a reversible hydrogen electrode (RHE). Linear sweep voltammetry (LSV) was tested at a scan rate of 5 mV/s. The Nyquist plots were measured by electrochemical impedance spectroscopy (EIS) in the frequency range from 100 kHz to 0.01 Hz with an amplitude of 10 mV. Cyclic voltammetry (CV) was used to measure the electric double-layer capacitance at non-faradaic potential and to estimate the effective electrode surface area with scan rates of 20, 40, 60, 100, 120, and 140 mV/s, respectively. The chronoamperometric measurements were used to evaluate the stabilities of the samples. All data were presented after IR compensation to reflect the actual catalytic currents.

4. Conclusions

Uniformly distributed CoS2 NWs are successfully prepared on SSF substrate by the hydrothermal synthesis of Co(OH)y(CO3)(1−0.5y)·nH2O NW precursors followed by thermal oxidation and a sulfuration process. After the thermal oxidation, the Co(OH)y(CO3)(1−0.5y)·nH2O NW precursors are converted into Co3O4 NWs with invisible morphology change. Based on experimental results, the catalytic activity of the CoS2/SSF electrode can be notably optimized by simply adjusting the sulfuration temperatures. The relatively high-purity CoS2 NWs with significantly bended morphology sulfured at 500 °C offer high conductivity and a large surface area for HER reactions. The overpotential to drive a current density of 100 mA·cm−2 is only 0.224 V, which is much lower than that of samples sulfured at lower temperatures.

Author Contributions

Conceptualization, H.-B.W.; Validation, H.-B.W. and L.Z.; Investigation, Z.-J.Q., H.Z. and H.-B.W.; Resources, H.-B.W.; Data curation, Z.-J.Q., H.-B.W. and H.Z.; Writing—original draft preparation, Z.-J.Q. and H.Z.; Writing—review and editing, H.-B.W.; Visualization, Z.-J.Q., H.Z. and H.-B.W.; Supervision, L.Z. and H.-B.W.; Project administration, D.-Y.M.; Funding acquisition, H.-B.W. and D.-Y.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work was jointly supported by the Natural Science Foundation of Shaanxi Province (No. 2020JM-216), the Key Research and Development project of Shandong Province (No. 2019GGX102023), and the foundation of State Key Laboratory for Mechanical Behavior of Mate-rials in XJTU (No. 20202204).

Data Availability Statement

All relevant data are contained in the present manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jing, S.; Zhang, L.; Luo, L.; Lu, J.; Yin, S.; Shen, P.K.; Tsiakaras, P. N-Doped Porous Molybdenum Carbide Nanobelts as Efficient Catalysts for Hydrogen Evolution Reaction. Appl. Catal. B Environ. 2018, 224, 533–540. [Google Scholar] [CrossRef]
  2. Darshan, B.N.; Kareem, A.; Maiyalagan, T.; Edwin Geo, V. CoS2/MoS2 decorated with nitrogen doped reduced graphene oxide and multiwalled carbon nanotube 3D hybrid as efficient electrocatalyst for hydrogen evolution reaction. Int. J. Hydrog. Energy 2021, 46, 13952–13959. [Google Scholar] [CrossRef]
  3. Ji, K.; Matras-Postolek, K.; Shi, R.; Chen, L.; Che, Q.; Wang, J.; Yue, Y.; Yang, P. MoS2/CoS2 heterostructures embedded in N-doped carbon nanosheets towards enhanced hydrogen evolution reaction. J. Alloys Compd. 2021, 891, 161962. [Google Scholar] [CrossRef]
  4. Vikraman, D.; Akbar, K.; Hussain, S.; Yoo, G.; Jang, J.Y.; Chun, S.H.; Jung, J.; Park, H.J. Direct synthesis of thickness-tunable MoS2 quantum dot thin layers: Optical, structural and electrical properties and their application to hydrogen evolution. Nano Energy 2017, 35, 101–114. [Google Scholar] [CrossRef]
  5. Hussain, S.; Akbar, K.; Vikraman, D.; Karuppasamy, K.; Kim, H.S.; Chun, S.H.; Jung, J. Synthesis of MoS2(1-x)Se2x and WS2(1-x)Se2x alloys for enhanced hydrogen evolution reaction performance. Inorg. Chem. Front. 2017, 4, 2068–2074. [Google Scholar] [CrossRef]
  6. Weng, B.; Wei, W.; Yu, Y.; Wu, H.; Al-Enizi, A.M.; Zheng, G. Bifunctional CoP and CoN porous nanocatalysts derived from ZIF-67 in situ grown on nanowire photoelectrodes for efficient photoelectrochemical water splitting and CO2 reduction. J. Mater. Chem. A 2016, 4, 15353–15360. [Google Scholar]
  7. Shang, X.; Yan, K.L.; Liu, Z.Z.; Lu, S.S.; Dong, B.; Chi, J.Q.; Li, X.; Liu, Y.R.; Chai, Y.M.; Liu, C.G. Oxidized carbon fiber supported vertical WS2 nanosheets arrays as efficient 3D nanostructure electrocatalyts for hydrogen evolution reaction. Appl. Surf. Sci. 2017, 402, 120–128. [Google Scholar] [CrossRef]
  8. Lv, Y.; Wang, X. Nonprecious metal phosphides as catalysts for hydrogen evolution, oxygen reduction and evolution reactions. Catal. Sci. Technol. 2017, 7, 3676–3691. [Google Scholar] [CrossRef]
  9. Sun, H.; Yan, Z.; Liu, F.; Xu, W.; Cheng, F.; Chen, J. Self-supported transition-metal-based electrocatalysts for hydrogen and oxygen evolution. Adv. Mater. 2020, 32, 1806326. [Google Scholar] [CrossRef] [PubMed]
  10. Yang, Z.; Zhao, H.; Dai, X.; Nie, F.; Ren, Z.; Yin, X.; Gan, Y.; Wu, B.; Cao, Y.; Zhang, X. Phosphorus-doping induced electronic modulation of CoS2–MoS2 hollow spheres on MoO2 film-Mo foil for synergistically boosting alkaline hydrogen evolution reaction. Int. J. Hydrogen Energy 2021, 46, 33388–33396. [Google Scholar] [CrossRef]
  11. Yang, J.; Zhang, C.; Niu, Y.; Huang, J.; Qian, X.; Wong, K.Y. N-doped C-CoS2@CoS2/MoS2 nano polyhedrons with hierarchical yolk-shelled structures as bifunctional catalysts for enhanced photovoltaics and hydrogen evolution. Chem. Eng. J. 2021, 409, 128293. [Google Scholar] [CrossRef]
  12. Zhou, Y.; Zhang, J.; Ren, H.; Pan, Y.; Yan, Y.; Sun, F.; Wang, X.; Wang, S.; Zhang, J. Mo doping induced metallic CoSe for enhanced electrocatalytic hydrogen evolution. Appl. Catal. B Environ. 2020, 268, 118467. [Google Scholar] [CrossRef]
  13. Lin, J.; He, J.; Qi, F.; Zheng, B.; Wang, X.; Yu, B.; Zhou, K.; Zhang, W.; Li, Y.; Chen, Y. In-situ Selenization of Co-based metal-organic frameworks as a highly efficient electrocatalyst for hydrogen evolution reaction. Electrochim. Acta 2017, 247, 258–264. [Google Scholar] [CrossRef]
  14. Ning, H.; Liu, Z.; Xie, Y.; Huang, H. CoS2 coatings for improving thermal stability and electrochemical performance of fes2 cathodes for thermal batteries. J. Electrochem. Soc. 2018, 165, A1725–A1733. [Google Scholar] [CrossRef]
  15. Zhu, Y.; Song, L.; Song, N.; Li, M.; Wang, C.; Lu, X. Bifunctional and efficient CoS2-C@MoS2 core-shell nanofiber electrocatalyst for water splitting. ACS Sustain. Chem. Eng. 2019, 7, 2899–2905. [Google Scholar] [CrossRef]
  16. Guo, Y.; Gan, L.; Shang, C.; Wang, E.; Wang, J. A Cake-Style CoS2@MoS2/RGO hybrid catalyst for efficient hydrogen evolution. Adv. Funct. Mater. 2017, 27, 1602699. [Google Scholar] [CrossRef]
  17. Shi, M.; Zhang, Y.; Zhu, Y.; Wang, W.; Wang, C.; Yu, A.; Pu, X.; Zhai, J. A flower-like CoS2/MoS2 heteronanosheet array as an active and stable electrocatalyst toward the hydrogen evolution reaction in alkaline media. RSC Adv. 2020, 10, 8973–8981. [Google Scholar] [CrossRef] [Green Version]
  18. Xie, W.; Liu, K.; Shi, G.; Fu, X.; Chen, X.; Fan, Z.; Liu, M.; Yuan, M.; Wang, M. CoS2 nanowires supported graphdiyne for highly efficient hydrogen evolution reaction. J. Energy. Chem. 2021, 60, 272–278. [Google Scholar] [CrossRef]
  19. Wang, F.; Zhao, J.; Tian, W.; Hu, Z.; Lv, X.; Zhang, H.; Yue, H.; Zhang, Y.; Ji, J.; Jiang, W. Morphology-controlled synthesis of CoMoO4 nanoarchitectures anchored on carbon cloth for high-efficiency oxygen oxidation reaction. RSC Adv. 2019, 9, 1562–1569. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Liang, Y.; Yao, S.; Wang, Y.; Yu, H.; Majeed, A.; Shen, X.; Li, T.; Qin, S. Hybrid cathode composed of pyrite-structure CoS2 hollow polyhedron and Ketjen black@sulfur materials propelling polysulfide conversion in lithium sulfur batteries. Ceram. Int. 2021, 47, 27122–27131. [Google Scholar] [CrossRef]
  21. Vijaya, S.; Landi, G.; Wu, J.J.; Anandan, S. Ni3S4/CoS2 mixed-phase nanocomposite as counter electrode for Pt-free dye-sensitized solar cells. J. Power Sour. 2020, 478, 229068. [Google Scholar] [CrossRef]
  22. Chen, Y.; Yang, W.; Gao, S.; Gao, Y.; Sun, C.; Li, Q. Catalytic reduction of aqueous bromate by a non-noble metal catalyst of CoS2 hollow spheres in drinking water at room temperature. Sep. Purif. Technol. 2020, 251, 117353. [Google Scholar] [CrossRef]
  23. Zhang, J.; Liu, T.; Fu, L.; Ye, G. Synthesis of nanosized ultrathin MoS2 on montmorillonite nanosheets by CVD method. Chem. Phys. Lett. 2021, 781, 138972. [Google Scholar] [CrossRef]
  24. Chen, F.; Yao, Y.; Su, W.; Zhao, S.; Ding, S.; Fu, L. The synthesis of 2D MoS2 flakes with tunable layer numbers via pulsed-Argon-flow assisted CVD approach. Ceram. Int. 2020, 46, 14523–14528. [Google Scholar] [CrossRef]
  25. Wu, Q.; Liu, L.; Guo, H.; Li, L.; Tai, X. Decorated by Cu nanoparticles CoS2 nanoneedle array for effective water oxidation electrocatalysis. J. Alloys Compd. 2020, 821, 153219. [Google Scholar] [CrossRef]
  26. Zhou, Y.; Zhang, W.; Zhang, M.; Shen, X.; Zhang, Z.; Meng, X.; Shen, X.; Zeng, X.; Zhou, M. Hetero-structured CoS2-MoS2 hollow microspheres with robust catalytic activity for alkaline hydrogen evolution. Appl. Surf. Sci. 2020, 527, 146847. [Google Scholar] [CrossRef]
  27. Yan, G.; Wu, C.; Tan, H.; Feng, X.; Yan, L.; Zang, H.; Li, Y. N-Carbon coated P-W2C composite as efficient electrocatalyst for hydrogen evolution reactions over the whole pH range. J. Mater. Chem. A 2017, 5, 765–772. [Google Scholar] [CrossRef]
  28. Liu, H.; Xia, G.; Zhang, R.; Jiang, P.; Chen, J.; Chen, Q. MOF-derived RuO2/Co3O4 heterojunctions as highly efficient bifunctional electrocatalysts for HER and OER in alkaline solutions. RSC Adv. 2017, 7, 3686–3694. [Google Scholar] [CrossRef] [Green Version]
  29. Ashassi-Sorkhabi, H.; Rezaei-Moghadam, B.; Asghari, E.; Bagheri, R.; Hosseinpour, Z. Fabrication of bridge like Pt@MWCNTs/CoS2 electrocatalyst on conductive polymer matrix for electrochemical hydrogen evolution. Chem. Eng. J. 2017, 308, 275–288. [Google Scholar] [CrossRef]
  30. Zhu, H.; Wang, H.B.; Zhu, R.H.; Zhou, L.; Cao, B.M.; Yu, Z.R.; Zeng, D.J.; Zhang, C.J.; Zhang, L.; Ma, D.Y. Ultra-thin CoS2 nanosheets synthesized by one-step hydrothermal process for hydrogen evolution. J. Mater. Sci. Mater. Electron. 2021, 32, 9149–9157. [Google Scholar] [CrossRef]
  31. 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 Lett. 2018, 3, 779–786. [Google Scholar] [CrossRef]
  32. Zhou, X.; Yang, X.; Li, H.; Hedhili, M.N.; Huang, K.W.; Li, L.J.; Zhang, W. Symmetric synergy of hybrid CoS2-WS2 electrocatalysts for the hydrogen evolution reaction. J. Mater. Chem. A 2017, 5, 15552–15558. [Google Scholar] [CrossRef] [Green Version]
  33. Liao, S.-Y.; Cui, T.T.; Zhang, S.Y.; Cai, J.J.; Zheng, F.; Liu, Y.D.; Min, Y.G. Cross-nanoflower CoS2 in-situ self-assembled on rGO sheet as advanced anode for lithium/sodium ion battery. Electrochim. Acta 2019, 326, 134992. [Google Scholar] [CrossRef]
  34. Zeng, S.; Tang, R.; Duan, S.; Li, L.; Liu, C.; Gu, X.; Wang, S.; Sun, D. Kinetically controlled synthesis of bismuth tungstate with different structures by a NH4F assisted hydrothermal method and surface-dependent photocatalytic properties. J. Colloid Interface Sci. 2014, 432, 236–245. [Google Scholar] [CrossRef]
  35. Turgut, G.; Sonmez, E.; Aydin, S.; Dilber, R.; Turgut, U. The effect of Mo and F double doping on structural, morphological, electrical and optical properties of spray deposited SnO2 thin films. Ceram. Int. 2014, 40, 12891–12898. [Google Scholar] [CrossRef]
  36. Xu, Y.F.; Ma, D.K.; Chen, X.A.; Yang, D.P.; Huang, S.M. Bisurfactant-controlled synthesis of three-dimensional YBO3/Eu3+ architectures with tunable wettability. Langmuir 2009, 25, 7103–7108. [Google Scholar] [CrossRef]
  37. Jun, S.E.; Choi, S.; Choi, S.; Lee, T.H.; Kim, C.; Yang, J.W.; Choe, W.O.; Im, I.H.; Kim, C.J.; Jang, H.W. Direct synthesis of molybdenum phosphide nanorods on silicon using graphene at the heterointerface for efficient photoelectrochemical water reduction. Nano-Micro Lett. 2021, 13, 135–150. [Google Scholar] [CrossRef]
  38. Qi, K.; Yu, S.; Wang, Q.; Zhang, W.; Fan, J.; Zheng, W.; Cui, X. Decoration of the inert basal plane of defect-rich MoS2 with Pd atoms for achieving Pt-similar HER activity. J. Mater. Chem. A 2016, 4, 4025–4031. [Google Scholar] [CrossRef]
Figure 1. The SEM images of (a) the SSF substate; (b) the Co(OH)y(CO3)(1−0.5y)·nH2O NWs precursors; (c) the CoO4 NWs; and the CoS2/SSF samples sulfureted at (df) 300 °C, (gi) 400 °C, and (jl) 500 °C.
Figure 1. The SEM images of (a) the SSF substate; (b) the Co(OH)y(CO3)(1−0.5y)·nH2O NWs precursors; (c) the CoO4 NWs; and the CoS2/SSF samples sulfureted at (df) 300 °C, (gi) 400 °C, and (jl) 500 °C.
Catalysts 12 00663 g001
Figure 2. (a) The XRD diffraction pattern of the untempered by heat treatment, the Co3O4 NWs samples, and CoS2/SSF; the XPS patterns of the CoS2 NWs: (b) the total profile; (c) Co 2p; and (d) the S 2p profile.
Figure 2. (a) The XRD diffraction pattern of the untempered by heat treatment, the Co3O4 NWs samples, and CoS2/SSF; the XPS patterns of the CoS2 NWs: (b) the total profile; (c) Co 2p; and (d) the S 2p profile.
Catalysts 12 00663 g002
Figure 3. (a) A TEM image of the CoS2 NWs; (b) a high-resolution TEM image; (c) the diffraction rings; (d) and an elemental distribution map (EDX).
Figure 3. (a) A TEM image of the CoS2 NWs; (b) a high-resolution TEM image; (c) the diffraction rings; (d) and an elemental distribution map (EDX).
Catalysts 12 00663 g003
Figure 4. CoS2/SSF at different vulcanization temperatures. (a) The LSV diagram; (b) the Tafel slope diagram; (c) the Nyquist diagram; and (d) the Cdl plots.
Figure 4. CoS2/SSF at different vulcanization temperatures. (a) The LSV diagram; (b) the Tafel slope diagram; (c) the Nyquist diagram; and (d) the Cdl plots.
Catalysts 12 00663 g004
Figure 5. The stability tests of the CoS2/SSF sample at sulfured 500 °C. (a) The chronoamperometric test result; (b) the polarization curves before and after 1000 cycles; (c,d) are the SEM images before and after the chronoamperometry testing.
Figure 5. The stability tests of the CoS2/SSF sample at sulfured 500 °C. (a) The chronoamperometric test result; (b) the polarization curves before and after 1000 cycles; (c,d) are the SEM images before and after the chronoamperometry testing.
Catalysts 12 00663 g005
Figure 6. A schematic diagram of the preparation process of the CoS2 NWs.
Figure 6. A schematic diagram of the preparation process of the CoS2 NWs.
Catalysts 12 00663 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, H.-B.; Qing, Z.-J.; Zhu, H.; Zhou, L.; Ma, D.-Y. Sulfuration Temperature-Dependent Hydrogen Evolution Performance of CoS2 Nanowires. Catalysts 2022, 12, 663. https://doi.org/10.3390/catal12060663

AMA Style

Wang H-B, Qing Z-J, Zhu H, Zhou L, Ma D-Y. Sulfuration Temperature-Dependent Hydrogen Evolution Performance of CoS2 Nanowires. Catalysts. 2022; 12(6):663. https://doi.org/10.3390/catal12060663

Chicago/Turabian Style

Wang, Hong-Bo, Zhuo-Jun Qing, Hao Zhu, Liang Zhou, and Da-Yan Ma. 2022. "Sulfuration Temperature-Dependent Hydrogen Evolution Performance of CoS2 Nanowires" Catalysts 12, no. 6: 663. https://doi.org/10.3390/catal12060663

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