Next Article in Journal
A Simplified Method for Assessing the Response of RC Frame Structures to Sudden Column Removal
Next Article in Special Issue
The Ethanol Oxidation Reaction Performance of Carbon-Supported PtRuRh Nanorods
Previous Article in Journal
Coordinated LVRT Control for a Permanent Magnet Synchronous Generator Wind Turbine with Energy Storage System
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Basal-Plane Catalytic Activity of Layered Metallic Transition Metal Ditellurides for the Hydrogen Evolution Reaction

1
Division of Chemical Engineering and Materials Science, Ewha Womans University, Seoul 03760, Korea
2
Department of Energy Science, Sungkyunkwan University, Suwon 16419, Korea
3
Korea Institute for Advanced Study, Seoul 02455, Korea
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Appl. Sci. 2020, 10(9), 3087; https://doi.org/10.3390/app10093087
Submission received: 17 March 2020 / Revised: 24 April 2020 / Accepted: 25 April 2020 / Published: 28 April 2020
(This article belongs to the Special Issue Nano Hydrogen Production and Storage)

Abstract

:
We report the electrochemical hydrogen evolution reaction (HER) of two-dimensional metallic transition metal dichalcogenides (TMDs). TMTe2 (TM: Mo, W, and V) single crystals were synthesized and characterized by optical microscopy, X-ray diffraction, and electrochemical measurements. We found that TMTe2 acts as a HER-active catalyst due to the inherent catalytic activity of its basal planes. Among the three metallic TMTe2, VTe2 shows the best HER performance with an overpotential of 441 mV and a Tafel slope of 70 mV/dec. It is 668 mV and 137 mV/dec for MoTe2 and 692 mV and 169 mV/dec for WTe2. Even though VTe2 has the lowest values in the exchange current density, the active site density, and turn-over-frequency (TOF) among the three TMTe2, the lowest charge transfer resistance (RCT) of VTe2 seems to be critical to achieving the best HER performance. First-principles calculations revealed that the basal-plane-active HER performance of metallic TMDs can be further enhanced with some Te vacancies. Our study paves the way to further study of the inherent catalytic activity of metallic 2D materials for active hydrogen production.

1. Introduction

Hydrogen energy is receiving tremendous attention as it is regarded as the most promising eco-friendly and renewable energy to replace the carbon emission energies such as fossil fuels [1,2,3,4]. To produce high purity hydrogen gas, electrochemical water splitting has been frequently used with noble metal-based electrocatalysts [5]. Much research has focused on the development of new types of the electrocatalysts, combining lower amounts of noble metals with earth-abundant materials for economy and efficient hydrogen production [6,7].
Hybrid electrocatalysts commonly consist of catalytic active nanoparticles embedded in catalytic inactive low-dimensional supports such as a graphene matrix with defects [8,9,10,11,12,13,14]. However, the high surface energy of nanoparticles has hindered the goal of obtaining small-sized nanoparticles and has boosted the aggregation of nanoparticles, which is known as the Ostwald ripening effect in nanoparticles [15,16]. Therefore, for a high number of active sites, the periodic arrangement of the catalytic active sites in the 2D plane can be a challenging theme for an efficient hydrogen evolution reaction (HER).
Two dimensional transition metal dichalcogenides (TMDs) are thought of as next generation electrocatalysts for the HER because of their abundant HER active sites with almost zero Gibbs free energy for hydrogen adsorption [17,18]. The semiconducting TMDs such as MoS2 and WS2 demonstrate excellent performance for hydrogen production at their active sites, mostly defects or edges [19,20,21]. Some TMDs experience a structural phase transition to a metallic state with defects, strains, or doping [22,23] and the thickness of TMDs has been known as an important property for the charge transfer process in HER performance [24]. Therefore, an unstable surface state due to many defects or edges is the main obstacle for stable hydrogen production.
The recently reported basal plane activity in metallic TMDs has been spotlighted to solve the above problems [18,23,24,25,26]. Soek et al. reported that the nonhomogeneous distribution of surface atoms creates intrinsic catalytic activity at its basal plane [22]. They also reported that the HER process of metallic TMDs sometimes accompanies structural modification due to hydrogen adsorption, which helps to enhance the catalytic activity of metallic TMDs. Kwon et al. also reported that the HER performance of metallic TMDs can be enhanced by the creation of Te vacancies [27]. Even though many metallic TMDs show the HER performance, the detailed mechanism of their basal plane-catalytic activity for HER is still unclear.
Here, we report the intrinsic catalytic activity of the basal plane of metallic transition metal (TM) ditellurides (TMTe2), where the TMs are V, Mo, and W. We synthesized TMTe2 single crystals using previously reported methods [27,28,29] and characterized it by optical microscopy, X-ray diffraction, and electrochemical measurements. X-ray diffraction results confirmed that TMTe2 single crystals have the orthorhombic, monoclinic, and monoclinic structure for WTe2, MoTe2, and VTe2, respectively [27,28,29]. The optical microscopy and the SEM images (Figure S1) show that their surfaces were well prepared with the clean surface minimizing the cracks, the edges, or the defects, which were usually regarded as extrinsic active sites in many TMDs. We found that clean surfaces of TMTe2 behaved as HER-active surfaces due to the inherent catalytic activity of their basal planes. Among the three metallic TMTe2, VTe2 achieved the best HER performance with the lowest overpotential and Tafel slope, which demonstrated that it was promising for active hydrogen production.

2. Experiments

X-ray diffraction of three TMTe2 single crystals was taken using an X-ray diffractometer (D/MAX-2500/PC, Rigaku, Japan) with Cu K radiation (λ = 1.54718 Å) at room temperature. The temperature dependent resistance of three TMTe2 single crystals was measured by a physical property measurement system (PPMS) (Quantum Design) through direct current (DC) measurements over a temperature range between 1.8 and 300 K.
For the electrochemical measurement of three TMTe2 single crystals, a piece of TMTe2 single crystal was glued to an Au (100 nm)/SiO2 (300 nm)/Si substrate and connected to the bottom Cu plate. Except for the open area for the HER, all other surface areas were covered by an insulating nail polish to prevent any unexpected leakage current. The open area of samples for the HER was automatically estimated by the Leopard 2.5 program installed in the software for the optical microscope; the current density (J) in this manuscript was calculated by the current divided by the open sample area.
All electrochemical experiments were performed using a potentiostat (VMP3 from Bio Logic, France) and three-electrode electrochemical cell in 0.5 M sulfuric acid (H2SO4) electrolyte solution. A silver/silver chloride electrode (Ag/AgCl, Qrins, Korea), a platinum wire electrode (Qrins, Korea), and the TMTe2 were used as the reference, counter, and working electrodes, respectively. Linear sweep voltammetry (LSV) was conducted with a scan rate of 5 mV/s. The electrochemical active surface area (ECSA) analysis with VTe2, MoTe2, and WTe2 was conducted at different voltage range and sweep rates. For impedance measurement, potentiostatic electrochemical impedance spectroscopy (PEIS) technique was used and measurements were carried out at frequency range between 100 kHz to 100 mHz with an AC voltage of 10 mV and a DC voltage, electric potential, of 390 mV (vs. RHE). The results were analyzed by using z-fitting software named EC-Lab.
First-principles calculations were conducted using a Quantum ESPRESSO package [30]. We used norm-conserving pseudopotentials [31] for W, Te, and H atoms, a plane wave energy cutoff of 80 Ryd, and a Perdew-Burke-Ernzerhof exchange-correlation functional [32]. Equilibrium lattice parameters of a = 6.36 Å and b = 3.53 Å are used to simulate the orthorhombic MoTe2 structure.

3. Result and Discussion

The atomic arrangements of the three metallic transition metal ditellurides (TMTe2), VTe2, WTe2, and MoTe2, are shown in Figure 1a. One transition metal atom is surrounded by six tellurium atoms with the form of a distorted octahedron: one center TM atom and six corner Te atoms. These layered metallic TMDs have exotic buckled layer structures due to the lattice distortion originating from strong correlations between spin, charge, and lattice in 2D materials [28,29]. The buckled monolayer of TMTe2 has non-uniform distribution of surface Te atoms, which is categorized as α (orange circles) and β (yellow circles) according to the vertical positions at the basal plane as seen in Figure 1a. Different from the zig-zag chains (black lines) of Mo or W atoms in MoTe2 and WTe2, VTe2 has mirrored double zig-zag chains (black lines) of V atoms, leading to the different surface densities of site α and β for hydrogen evolution as described in Table 1.
Different Te atoms in the buckled layer of the three TMTe2 have different Gibbs free energies for hydrogen adsorption. We classified two Te atomic sites as α sites and β sites to analyze the HER performances of metallic TMTe2. We assume that α site Te atoms are better catalytically active than β site Te atoms in the three TMTe2, which is consistent with recently reported HER works on MoTe2 and WTe2 [22,23,27]. During the HER process, approaching proton atoms (H+) are captured at α sites rather than β sites due to the differences in Gibbs free energies, which creates the intrinsic catalytic activity for hydrogen evolution [22,23,27]. The number densities of active α site in three TMTe2 are calculated with unit surface area, which contains one active α site (squares with dashed lines in Figure 1a) in the basal plane (see Table 1).
To study the inherent catalytic activity of the basal plane in TMTe2, TMTe2 single crystals were used for this study. We synthesized single crystalline TMTe2 using the NaCl-Flux method and the solid state reaction described in our previous studies [28,29]. We confirmed that TMTe2 single crystals have the orthorhombic (a = 3.498 Å, b = 6.338 Å, c = 15.432 Å, β = 90°), monoclinic (a = b = 3.559 Å, c = 40.693 Å, β = 90°), and monoclinic (a = 14.826 Å, b = 3.605 Å, c= 9.366 Å, β = 112.4°) structure for WTe2, MoTe2, and VTe2, respectively, and the TMTe2 was well grown with a growth axis in the (00n) direction using x-ray diffraction as seen in Figure 1b. All three TMTe2 show metallic behaviors in which the resistivity decreases with a decreasing temperature (see Figure S2). The three TMTe2 are known as semimetals (or bad metals) with few carriers due to the quasi gap at the Fermi surface [33,34]. Thus, they have good charge transfer and also less of a screening effect.
Figure 2a shows the experimental setup for electrochemical measurements of the three metallic TMTe2. We used a three-electrode system for the TMTe2 working electrode (WE), platinum wire for the counter electrode (CE), and Ag/AgCl for the reference electrode (RE). We applied an external voltage between the WE and RE and measured the current density between the WE and CE. TMTe2 single crystals were used for the WE and the sample area of the WE was estimated by an analyzing program (Leopard 2.5) with optical images.
The electrochemical HER performance of the three TMTe2 electrodes was examined by linear sweep voltammetry (LSV) and electrochemical impedance spectroscopy (EIS). Figure 2b shows that the LSV curves of three TMTe2 electrodes and the overpotentials of −441, −668, and −692 mV were required to obtain 10 mA/cm2 for VTe2, MoTe2, and WTe2, respectively. The overpotential of VTe2 was found to be the smallest among three TMTe2 electrodes. Figure 2c shows the Tafel slopes of the three TMTe2 electrodes were 70, 137, and 169 mV/dec for VTe2, MoTe2, and WTe2, respectively, and VTe2 exhibited the lowest Tafel slope among them. The overpotential and the Tafel slope of VTe2 single crystal were consistent with reported values in micro-sized VTe2 flakes, which were transferred to the glassy carbon working electrode by drop-cast method, indicating that the basal plane of VTe2 can be active for HER [35]. We extracted the exchange current densities of three TMTe2 from extended linear lines of Tafel slopes down to zero value in the overpotential, 0.2, 0.7, and 1.0 μA/cm2 for VTe2, MoTe2, and WTe2, respectively.
To analyze the catalytic activity of the α site in the TMTe2, we calculated the turn over frequency (TOF), which is proportional to the conversion rate from proton to hydrogen gas at one active site for one second. The TOF was calculated by equation TOF = j × A × N A S × n × F , where j is the exchange current density, n is the number of electrons transferred per molecule (2 for HER), F is the Faraday constant (96,458 C mol−1 electrons), A is the superficial electrode area, and NA is Avogadro’s number (6.022 × 1023 mol−1). The TOF values were 2.14, 4.85, and 6.92 ms−1 for VTe2, MoTe2, and WTe2, respectively, and WTe2 had the largest TOF value among the three. It seems that the TOF results contradict the reported Gibbs free energies of α sites in MoTe2 (ΔGH = 0.77 eV) and WTe2 (ΔGH = 1.14 eV) [17,18]. This is explained by the fact that the hydrogen adsorption at the catalytic active sites is less critical in determining the hydrogen conversion rate during the HER process.
To study the charge transfer process for the HER, we measured the electrochemical impedance spectroscopy (EIS) with the three TMTe2 electrodes. Figure 2d shows the Nyquist plots of the three TMTe2 electrodes with the complex impedance, the real part ZRe, and the imaginary part −ZIm with various frequencies from 100 kHz to 100 mHz. We included the Warburg impedance (w) on the equivalent Randles circuit models to interpret the mass transfer at the low frequency region (inset in Figure 2d). The Nyquist plots show semicircles in the kinetic control region at the high frequencies with the diameter of RCT, which represents the charge transfer resistance. The values of RCT were 5.8, 13.0, and 21.6 kΩ, which increased in order of VTe2, MoTe2, and WTe2 and the lowest RCT was found in VTe2.
We also estimated the double layer capacitance (CDL) between the TMTe2 electrode and the electrolyte. Cyclic voltammetry (CV) was measured with different scan rates. Figure 3a shows the hysteresis of the CV curves taken in non-Faradaic voltage regions without any additional electrochemical reactions. We plotted the curves of the capacitive current density as shown in the CV curves in Figure 3a as a function of scan rate to obtain CDL of three TMTe2 electrodes. Figure 3b shows similar CDL values due to the similar sample area, the same dielectric constant of the electrolyte, and the same distance of the double layer capacitance. These CDL values in TMTe2 single crystals were almost 100 times smaller than the reported CDL in micro-sized flakes on the glassy carbon working electrode due to the difference of the electrochemical surface area [23].
We summarized our experimental results of the HER performances observed in the three TMTe2 in Table 1. Among the three TMTe2, VTe2 had the best HER performance with a small overpotential and Tafel slope. Even though VTe2 had the smallest α site density and TOF, the smallest RCT was found in VTe2. It seems that the HER process of TMTe2 is significantly influenced by the charge transfer process. Therefore, to enhance the inherent catalytic activity of TMTe2, we can further enhance the charge transfer for a small RCT or to increase the electrochemical active surface area for large CDL.
Considering the Gibbs free energies of α sites in TMTe2, the defect creation can help to enhance the hydrogen adsorption on basal plane of TMTe2. Figure 4 shows the energies of hydrogen adsorption of α and β sites in MoTe2 with or without Te defects. We can decrease the energy of hydrogen adsorption at α sites from 0.84 eV to 0.21 eV with Te defects, which can be a considerable value for better HER.

4. Conclusions

In conclusion, the basal plane catalytic activity for the HER was investigated with three metallic TMTe2 (TM: Mo, W, and V). We measured the electrochemical measurements, LSVs, EISs, and CVs, and obtained the overpotential, Tafel slope, exchange current, α site density, TOF, CDL, and RCT. Among the three TMTe2, VTe2 had the best HER performance with an overpotential of −441 mV and a Tafel slope of 70 mV/dec. Even though VTe2 had the lowest exchange current, the α site density and TOF among the three TMTe2, the lowest RCT of VTe2 seems to be critical to achieve the best HER performance. Our study suggests that the inherent HER performance of metallic TMDs can be enhanced by enhancing the charge transfer process, increasing the electrochemical active surface area or by creating Te vacancies.

Supplementary Materials

The following are available online at https://www.mdpi.com/2076-3417/10/9/3087/s1, Figure S1: Optical image and SEM image of WTe2 single crystal for the electrochemical measurements, Figure S2: Residual resistance ratio (RRR) of three TMTe2 single crystals.

Author Contributions

H.K.: Data curation, Investigation, Writing —Original draft, Writing—review & editing. D.B.: Data curation, Investigation, Writing—Original draft. H.J.: Data curation, Investigation. B.J.: Investigation, D.W.: Investigation. J.-H.L.: Formal analysis. Y.-W.S.: Formal analysis. H.Y.: Conceptualization, Supervision. S.C.: Conceptualization, Supervision, Writing—Original draft, Writing - review & editing. All authors have read and agreed to the published version of the manuscript.

Acknowledgments

S.C.: H.K. and D.B. were supported by the Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Science, ICT, and Future Planning (2020R1A2C2003377). H.K. and H. J. were supported by the National Research Foundation of Korea (NRF) and Center for Women In Science, Engineering and Technology (WISET) grant funded by the Ministry of Science and ICT (MSIT) under the team research program for female engineering students. H.Y. acknowledges support from the National Research Foundation of Korea (NRF) under Grant No. NRF-2020R1A2B5B02002548. Y.-W.S. was supported by the NRF of Korea (Grant No. 2017R1A5A1014862, SRC program: vdWMRC Center). We thank the Korea Institute for Advanced Study for providing computing resources (KIAS Center for Advanced Computation Linux Cluster System) for this work.

Conflicts of Interest

The authors declare no competing financial interests.

References

  1. Dong, G.; Fang, M.; Wang, H.; Yip, S.; Cheung, H.Y.; Wang, F.; Wong, C.Y.; Chu, S.T.; Ho, J.C. Insight into the electrochemical activation of carbon-based cathodes for hydrogen evolution reaction. J. Mater. Chem. A 2015, 3, 13080. [Google Scholar] [CrossRef]
  2. Zhang, L.; Han, L.; Liu, H.; Liu, X.; Luo, J. Potential-cycling synthesis of single platinum atoms for efficient hydrogen evolution in neutral media. Angew. Chem. 2017, 129, 13882. [Google Scholar] [CrossRef] [Green Version]
  3. Sun, S.W.; Wang, G.F.; Zhou, Y.; Wang, F.B.; Xia, X.H. High-performance Ru@C4N electrocatalyst for hydrogen evolution reaction in both acidic and alkaline solutions. ACS Appl. Mater. Interfaces 2019, 11, 19176–19182. [Google Scholar] [CrossRef] [PubMed]
  4. Wang, Y.; Sofer, Z.; Luxa, J.; Pumera, M. Lithium exfoliated vanadium dichalcogenides (VS2, VSe2, VTe2) exhibit dramatically different properties from their bulk counterparts. Adv. Mater. Interfaces 2016, 3, 1600433. [Google Scholar] [CrossRef]
  5. Li, C.; Baek, J.B. Recent advances in noble metal (Pt, Ru, and Ir)-based electrocatalysts for efficient hydrogen evolution reaction. ACS Omega 2020, 5, 31–40. [Google Scholar] [CrossRef] [Green Version]
  6. Tavakkoli, M.; Holmberg, N.; Kronberg, R.; Jiang, H.; Saino, J.; Kauppinen, E.I.; Kallio, T.; Lassonen, K. Electrochemical activation of single-walled carbon nanotubes with Pseudo-atomic-scale platinum for the hydrogen evolution reaction. ACS Catal. 2017, 7, 3121. [Google Scholar] [CrossRef]
  7. Chen, J.; Yang, Y.; Su, J.; Jiang, P.; Zia, G.; Chen, Q. Enhanced activity for hydrogen evolution reaction over CoFe catalysts by alloying with small amount of Pt. ACS Appl. Mater. Interfaces 2017, 9, 4. [Google Scholar] [CrossRef]
  8. Jiang, B.; Sun, Y.; Cheng, Y.; Shao, M. Pt nanocrystals on nitrogen-doped graphene for the hydrogen evolution reaction using Si nanowires as a sacrificial template. Nanoscale 2017, 9, 10138–10144. [Google Scholar] [CrossRef]
  9. Liu, D.; Li, L.; You, T. Superior catalytic performances of platinum nanoparticles loaded nitrogen-doped graphene toward methanol oxidation and hydrogen evolution reaction. J. Colloid Interface Sci. 2017, 487, 330–335. [Google Scholar] [CrossRef]
  10. Li, X.; Zhang, L.; Zang, X.; Li, X.; Zhu, H. Photo-promoted platinum nanoparticles decorated MoS2@Graphene woven fabric catalyst for efficient hydrogen generation. ACS Appl. Mater. Interfaces 2016, 8, 10866–10873. [Google Scholar] [CrossRef]
  11. Li, W.; Amiinu, I.S.; Ye, B.; Wang, Z.; Zhu, J.; Kou, Z.; Mu, S. TePtFe nanotubes as high-performing bifunctional electrocatalysts for the oxygen reduction reaction and hydrogen evolution reaction. ChemSusChem 2018, 11, 1328–1333. [Google Scholar] [CrossRef] [PubMed]
  12. Sugawara, Y.; Okayasu, T.; Yadav, A.P.; Nishikata, A.; Tsuru, T. Dissolution mechanism of platinum in sulfuric acid solution. J. Electrochem. Soc. 2012, 159, F779–F786. [Google Scholar] [CrossRef]
  13. Seok, J.; Lee, J.; Bae, D.; Ji, B.; Son, Y.; Lee, Y.H.; Yang, H.; Cho, S. Hybrid catalyst with monoclinic MoTe2 and platinum for efficient hydrogen evolution. APL Mater. 2019, 7, 071118. [Google Scholar] [CrossRef] [Green Version]
  14. Ling, N.; Wang, Z.; Kim, S.; Oh, S.H.; Park, J.H.; Shin, H.; Cho, S.; Yang, H. In operando stacking of reduced graphene oxide for active hydrogen evolution. ACS Appl. Mater. Interfaces 2019, 11, 43460–43465. [Google Scholar] [CrossRef] [PubMed]
  15. Hansen, T.W.; DeLaRiva, A.T.; Challa, S.R.; Datye, A.K. Sintering of catalytic nanoparticles: Particle migration or Ostwald rippening? Acc. Chem. Res. 2013, 46, 1720–1730. [Google Scholar] [CrossRef] [PubMed]
  16. Jones, J.; Xiong, H.; De La Riva, A.T.; Peterson, E.J.; Pham, H.; Challa, S.R.; Qi, G.; Oh, S.; Wiebenga, M.H.; Hernández, X.I.P.; et al. Thermally stable single-atom platinum-on-ceria catalysts via atom trapping. Science 2016, 353, 150–154. [Google Scholar] [CrossRef] [Green Version]
  17. Hinnemann, B.; Moses, P.G.; Bonde, J.; Jørgensen, K.P.; Nielsen, J.H.; Horch, S.; Chorkendorff, I.; Nørskov, J.K. Biomimetic hydrogen evolution:  MoS2 nanoparticles as catalyst for hydrogen evolution. J. Am. Chem. Soc. 2005, 127, 5308–5309. [Google Scholar] [CrossRef]
  18. Lee, J.; Kang, S.; Yim, K.; Kim, K.Y.; Jang, H.W.; Kang, Y.; Han, S. Hydrogen evolution reaction at anion vacancy of two-dimensional transition-metal dichalcogenides: Ab initio computational screening. J. Phys. Chem. Lett. 2018, 9, 2049–2055. [Google Scholar] [CrossRef]
  19. Li, G.; Zhang, D.; Qiao, Q.; Yu, Y.; Peterson, D.; Zafar, A.; Kumar, R.; Curtarolo, S.; Hunte, F.; Shannon, S.; et al. All the catalytic active sites of MoS2 for hydrogen evolution. J. Am. Chem. Soc. 2016, 138, 16632. [Google Scholar] [CrossRef]
  20. Tsai, C.; Chan, K.; Abild-Pedersen, F.; Nørskov, J.K. Active edge sites in MoSe2 and WSe2 catalysts for the hydrogen evolution reaction: A density functional study. Phys. Chem. Chem. Phys. 2014, 16, 13156–13164. [Google Scholar] [CrossRef]
  21. Xie, J.; Zhang, H.; Li, S.; Wang, R.; Sun, X.; Zhou, M.; Zhou, J.; Lou, X.W.; Xie, Y. Defect-rich MoS2 ultrathin nanosheets with additional active edge sites for enhanced electrocatalytic hydrogen evolution. Adv. Mater. 2013, 25, 5807–5813. [Google Scholar] [CrossRef] [PubMed]
  22. Seok, J.; Lee, J.; Cho, S.; Ji, B.; Kim, H.W.; Kwon, M.; Kim, D.; Kim, Y.; Oh, S.H.; Kim, S.W.; et al. Active hydrogen evolution through lattice distortion in metallic MoTe2. 2D Mater. 2017, 4, 025061. [Google Scholar] [CrossRef] [Green Version]
  23. McGlynn, J.C.; Dankwort, T.; Kienle, L.; Bandeira, N.A.G.; Fraser, J.P.; Gibson, E.K.; Cascallana-Matías, I.; Kamarás, K.; Symes., M.D.; Miras, H.N.; et al. The rapid electrochemical activation of MoTe2 for the hydrogen evolution reaction. Nat. Commun. 2019, 10, 4916. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Mc Manus, J.B.; Cunningham, G.; McEvoy, N.; Cullen, C.P.; Gity, F.L.; Schmidt, M.; Mullarkey, D.; Shvets, I.V.; Hurley, P.K.; Hallam, T.; et al. Growth of 1T’ MoTe2 by thermally assisted conversion of electrodeposited tellurium films. ACS Appl. Energy Mater. 2019, 2, 521–530. [Google Scholar] [CrossRef]
  25. Rajamathi, C.R.; Gupta, U.; Kumar, N.; Yang, H.; Sun, Y.; Süß, V.; Shekhar, C.; Schmidt, M.; Blumtritt, H.; Werner, P.; et al. Weyl semimetal as hydrogen evolution catalyst. Adv. Mater. 2017, 29, 1606202. [Google Scholar] [CrossRef]
  26. Luxa, J.; Vosecký, P.; Mazánek, V.; Sedmidubský, D.; Pumera, M.; Lazar, P.; Sofer, Z. Layered transition-metal ditellurides in electrocatalystic applications—Contrasting properties. ACS Catal. 2017, 7, 5706–5716. [Google Scholar] [CrossRef]
  27. Kwon, H.; Ji, B.; Bae, D.; Lee, J.H.; Park, H.J.; Kim, D.H.; Kim, Y.M.; Son, Y.W.; Yang, H.; Cho, S. Role of anionic vacancy for active hydrogen evolution in WTe2. Appl. Surf. Sci. 2020, 515, 145972. [Google Scholar] [CrossRef]
  28. Keum, D.H.; Cho, S.; Kim, J.H.; Choe, D.; Sung, H.; Kan, M.; Kang, H.; Hwang, J.; Kim, S.W.; Yang, H.; et al. Bandgap opening in few-layered monoclinic MoTe2. Nat. Phys. 2015, 11, 482–486. [Google Scholar] [CrossRef]
  29. Won, D.; Kiem, D.H.; Cho, H.; Kim, D.; Kim, Y.; Jeong, M.Y.; Seo, C.; Kim, J.; Park, J.; Han, M.J.; et al. Polymorphic spin, charge, and lattice waves in vanadium ditelluride. Adv. Mater. 2020, 1906578, 1–7. [Google Scholar] [CrossRef]
  30. Giannozzi, P.; Baroni, S.; Bonini, N.; Calandra, M.; Car, R.; Cavazzoni, C.; Ceresoli, D.; Chiarotti, G.L.; Cococcioni, M.; Dabo, I.; et al. QUANTUM ESPRESSO: A modular and open-source software project for quantum simulations of materials. J. Phys. Condens. Matter. 2009, 21, 395502. [Google Scholar] [CrossRef]
  31. Hamann, D.R.; Schlüter, M.; Chiang, C. Norm-conserving pseudopotentials. Phys. Rev. Let. 1979, 43, 1494. [Google Scholar] [CrossRef]
  32. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Let. 1996, 77, 3865. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Cho, S.; Kang, S.H.; Yu, H.S.; Kim, H.W.; Ko, W.; Hwang, S.W.; Han, W.H.; Choe, D.H.; Jung, Y.H.; Chang, K.J.; et al. Te vacancy-driven superconductivity in orthorhombic molybdenum ditelluride. 2D Mater. 2017, 4, 021030. [Google Scholar] [CrossRef]
  34. Ali, M.N.; Xiong, J.; Flynn, S.; Tao, J.; Gibson, Q.D.; Schoop, L.M.; Liang, T.; Haldolaarachchige, N.; Hirschberger, M.; Ong, N.P.; et al. Large, non-saturating magnetoresistance in WTe2. Nature 2014, 514, 205–208. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Chia, X.; Ambrosi, A.; Lazar, P.; Sofer, Z.; Pumera, M. Electrocatalysis of layered group 5 metallic transition metal dichalcogenides (MX2, M= V, Nb, and Ta; X= S, Se, and Te). J. Mater. Chem. A 2016, 4, 14241. [Google Scholar] [CrossRef]
Figure 1. (a) Schematics of monolayer TMTe2 (TM = V, W, and Mo) with side and top view. Protons (H+) in acidic electrolyte (0.5 M H2SO4) are just passing without reactions at catalytic inactive Te atoms (β site, yellow circles), but are adsorbed at catalytic active Te atoms (α site, orange circles) and subsequently are converted to hydrogen gas (H2). (WTe2; a = 3.498 Å and b = 6.338 Å, MoTe2; a = 3.487 Å and b = 6.366 Å, VTe2; b = 3.605 Å, a+2c = 9.529 Å). (b) X-ray diffraction of three TMTe2 single crystals.
Figure 1. (a) Schematics of monolayer TMTe2 (TM = V, W, and Mo) with side and top view. Protons (H+) in acidic electrolyte (0.5 M H2SO4) are just passing without reactions at catalytic inactive Te atoms (β site, yellow circles), but are adsorbed at catalytic active Te atoms (α site, orange circles) and subsequently are converted to hydrogen gas (H2). (WTe2; a = 3.498 Å and b = 6.338 Å, MoTe2; a = 3.487 Å and b = 6.366 Å, VTe2; b = 3.605 Å, a+2c = 9.529 Å). (b) X-ray diffraction of three TMTe2 single crystals.
Applsci 10 03087 g001
Figure 2. (a) Experimental set-up for electrochemical measurements of three TMTe2 single crystals. (b) Linear sweep voltammetry and (c) Tafel plots and (d) the electrochemical impedance spectroscopy of three TMTe2 with the Randles circuit, which include the series resistance (RS), the charge transfer resistance (RCT), and the constant phase element (CPE).
Figure 2. (a) Experimental set-up for electrochemical measurements of three TMTe2 single crystals. (b) Linear sweep voltammetry and (c) Tafel plots and (d) the electrochemical impedance spectroscopy of three TMTe2 with the Randles circuit, which include the series resistance (RS), the charge transfer resistance (RCT), and the constant phase element (CPE).
Applsci 10 03087 g002
Figure 3. (a) Cyclic voltammetry (CV) curves at different scan rates and (b) capacitive current as a function of scan rate of VTe2, MoTe2, and WTe2.
Figure 3. (a) Cyclic voltammetry (CV) curves at different scan rates and (b) capacitive current as a function of scan rate of VTe2, MoTe2, and WTe2.
Applsci 10 03087 g003
Figure 4. Hydrogen adsorption energies at α and β sites of TMTe2 with or without Te vacancies.
Figure 4. Hydrogen adsorption energies at α and β sites of TMTe2 with or without Te vacancies.
Applsci 10 03087 g004
Table 1. Summary of electrochemical properties of three TMTe2.
Table 1. Summary of electrochemical properties of three TMTe2.
CatalystOverpotential
(mV vs. RHE)
Tafel Slope
(mV/dec)
Exchange Current Density
(mA/cm2)
⚜ site density
(#/nm2)
TOF
(ms−)
Double Layer Capacitance
(mF/cm2)
Charge Transfer Resistance
(k⚜CT)
VTe2−441700.22.922.140.035.8
MoTe2−6681370.74.504.850.0813.0
WTe2−6921691.04.516.920.0621.6

Share and Cite

MDPI and ACS Style

Kwon, H.; Bae, D.; Jun, H.; Ji, B.; Won, D.; Lee, J.-H.; Son, Y.-W.; Yang, H.; Cho, S. Basal-Plane Catalytic Activity of Layered Metallic Transition Metal Ditellurides for the Hydrogen Evolution Reaction. Appl. Sci. 2020, 10, 3087. https://doi.org/10.3390/app10093087

AMA Style

Kwon H, Bae D, Jun H, Ji B, Won D, Lee J-H, Son Y-W, Yang H, Cho S. Basal-Plane Catalytic Activity of Layered Metallic Transition Metal Ditellurides for the Hydrogen Evolution Reaction. Applied Sciences. 2020; 10(9):3087. https://doi.org/10.3390/app10093087

Chicago/Turabian Style

Kwon, Hagyeong, Dongyeon Bae, Hyeyoung Jun, Byungdo Ji, Dongyeun Won, Jun-Ho Lee, Young-Woo Son, Heejun Yang, and Suyeon Cho. 2020. "Basal-Plane Catalytic Activity of Layered Metallic Transition Metal Ditellurides for the Hydrogen Evolution Reaction" Applied Sciences 10, no. 9: 3087. https://doi.org/10.3390/app10093087

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