Skip to main content
ACS Omega logoLink to ACS Omega
. 2022 Feb 24;7(9):7837–7844. doi: 10.1021/acsomega.1c06730

Theoretical Insights into the Hydrogen Evolution Reaction on VGe2N4 and NbGe2N4 Monolayers

Mihir Ranjan Sahoo , Avijeet Ray , Nirpendra Singh §,∥,*
PMCID: PMC8908508  PMID: 35284711

Abstract

graphic file with name ao1c06730_0008.jpg

Catalytically active sites at the basal plane of two-dimensional monolayers for hydrogen evolution reaction (HER) are important for the mass production of hydrogen. The structural, electronic, and catalytic properties of two-dimensional VGe2N4 and NbGe2N4 monolayers are demonstrated using the first-principles calculations. The dynamical stability is confirmed through phonon calculations, followed by computation of the electronic structure employing the hybrid functional HSE06 and PBE+U. Here, we introduced two strategies, strain and doping, to tune their catalytic properties toward HER. Our results show that the HER activity of VGe2N4 and NbGe2N4 monolayers are sensitive to the applied strain. A 3% tensile strain results in the adsorption Gibbs free energy (ΔGH*) of hydrogen for the NbGe2N4 monolayer of 0.015 eV, indicating better activity than Pt (−0.09 eV). At the compressive strain of 3%, the ΔGH* value is −0.09 eV for the VGe2N4 monolayer, which is comparable to that of Pt. The exchange current density for the P doping at the N site of the NbGe2N4 monolayer makes it a promising electrocatalyst for HER (ΔGH* = 0.11 eV). Our findings imply the great potential of the VGe2N4 and NbGe2N4 monolayers as electrocatalysts for HER activity.

Introduction

The search for new catalytic materials to fulfill the demand for substantial energy resources and the development of alternative, efficient, cost-effective, abundant energy sources have received significant attention within the scientific community in the last few decades. The electrochemical water splitting through the hydrogen evolution reaction (HER) is an efficient method to produce hydrogen, an eco-friendly fuel with a high energy density, that is considered a promising technique to store energy from sustainable sources.13 The well-known electrocatalysts for HER are mostly noble metals like Pt and Pd and their compounds.46 However, high cost and low abundance hinder their applications in the mass production of hydrogen.7

In the last few decades, two-dimensional (2D) materials have been considered promising catalysts for HER because of their large surface-to-volume ratio and more active sites. Apart from that, 2D materials are excellent substrates due to their unique structural and electronic properties. The catalytic activity can be regulated through chemical modification and structural engineering.8 In addition, low cost, earth abundance, and ability to form various nanostructures have led to 2D materials gaining significant research attention in the field of catalysis911 and being considered as alternatives to the widely used and costly Pt-based catalysts.1214 The major drawback of 2D catalysts is that the catalytically active sites are confined to only edges while the basal plane is inert.1517 Although previous experimental and theoretical studies reported that the basal planes of MXenes are catalytically active sites for HER,1820 their performances are inferior to those of Pt and Pt-based compounds.21,22 Recently, graphene-based materials,23,24 reduced graphene oxides,25,26 g-C3N4,27 borophene,28 phosphorene,29 monolayers of Mo2C,30 and h-B2O31 have been proposed as promising catalysts. Furthermore, the HER performance of these 2D materials can be triggered through various strategies, such as surface functionalization,3234 intrinsic defect,35,36 doping,3739 and strain engineering.40 Despite numerous attempts to develop efficient 2D HER catalysts, the promising materials that can be suitable for practical device applications and take the place of Pt are still far from reality. Hence, the quest for efficient 2D electrocatalysts is essential for hydrogen energy.

Since a new 2D van der Waals (2D vdW) MA2Z4 family is recently synthesized using the chemical vapor deposition method,41 the attention on these 2D materials is rapidly increasing due to their potential applications.4246 The high carrier mobility and excellent stability make them suitable substrates. Moreover, hydrogen adsorption on the basal plane can be tuned due to the presence of the lowest unoccupied energy level, which leads to promising electrocatalysts for HER.47 The HER performance of 2D MoSi2N4 and WSi2N4 can be enhanced by introducing N vacancy and doping by transition metal atoms like V, Fe, Nb, Tc, and Ta.48 In addition, O doping at the N site and P, Fe, and Nb doping at the Si site of 2D MoSi2N4 exhibit superior HER activity.49 Moreover, an intercalated architecture approach is employed to predict 70 family members that are both dynamically and thermodynamically stable.50 The recent HER studies on the MA2Z4 class of material have been mainly focused on MoSi2N4 and WSi2N4 monolayers only, whereas other possible 2D magnetic materials of this family have not been explored. Recently, a multilevel screening of the 2D MA2Z4 family is performed by Liu et al.47 to explore the basal plane active catalyst for hydrogen evolution reaction; among 144 materials, they showed that compounds based on V and Nb metals with Ge and N show better performance by considering lowest occupied state energy, which ranges between −6.0 and −5.6 eV for better HER activity, as a descriptor. However, the catalytic activities of these materials, with respect to structural change by physical or chemical means, have not been discussed so far. Therefore, exploring the novel electronic and catalytic properties of other possible 2D magnetic materials (having more catalytically active sites in a basal plane) of this family with strain, defects, doping, etc. is of great interest for the fundamental research and practical applications in the field of hydrogen evolution reaction.

We present a systematic study on the structural, electronic, and catalytic properties of monolayer VGe2N4 and NbGe2N4 using the first-principles calculations. The dynamical stability is confirmed through the phonon band structure. The calculated Gibbs free energy (ΔGH*) is −0.29 eV (VGe2N4) and 0.19 eV (NbGe2N4), which are close to the ideal value (for Pt, ΔGH* = −0.09 eV51). The defects (N vacancy or Ge vacancy) and doping of nonmetal atoms (B, C, O, P, and S) are introduced to examine the performance of HER activity. Our findings will guide experimentalists in designing magnetic VGe2N4 and NbGe2N4 monolayers as excellent electrocatalysts for hydrogen evolution reactions.

Computational Methodology

The first-principles calculations are performed using density functional theory as employed in the Vienna ab initio simulation package (VASP).52,53 The exchange–correlation functional prescribed by Perdew, Burke, and Ernzerhof (PBE)54 under the generalized gradient approximation is used for the calculation of structural and electronic properties. The spin-polarized calculations of 2 × 2 × 1 supercell are performed with a Γ-centered 6 × 6 × 1 k-grids to sample the Brillouin zone. The kinetic energy cutoff of 500 eV is used to describe the plane-wave basis set. The monolayers are designed to maintain periodicity along the XY-plane with a vacuum of 20 Å along the z-direction to avoid the interactions between the periodic artificial images. Full relaxation of all of the atoms is allowed until a residual force of 0.001 eV/Å is reached. The Heyd, Scuseria, and Ernezerhof screened hybrid density functional (HSE06)55 with van der Waals correction is used to estimate the accurate band gaps of VGe2N4 and NbGe2N4 monolayers. We also performed PBE+U functional to account for the onsite Coulomb interaction of d electrons of V and Nb atoms. The finite displacement method56 is employed to calculate the phonon band structure.

The HER performance of VGe2N4 and NbGe2N4 monolayers is described by calculating the Gibbs free energy of the reaction intermediate (H*), which is defined as

graphic file with name ao1c06730_m001.jpg 1

where ΔEZPE and TΔSH* are the change in zero-point energy and entropy between atomic hydrogen adsorption and hydrogen in the gas phase, respectively. The contributions from the catalyst to both ΔEZPE and TΔSH* are very small and hence can be neglected. Thus, ΔEZPE can be obtained through the following equation57

graphic file with name ao1c06730_m002.jpg 2

where EZPEnH represents the zero-point energy of n hydrogen atoms adsorbed on monolayer without the contribution of the catalyst and EZPE is the zero-point energy of H2 molecule in the gas phase. The value of ΔEZPE can vary from −0.01 to 0.04 eV. The entropy of atomic hydrogen ΔSH (≈ −1/2 ΔSH0), where ΔSH is the entropy of H2 molecule in the gas phase. The value of TSH0 can be considered as 0.4 eV51 at T = 300 K, which gives rise to the value of ΔEZPETΔSH* equal to 0.24 eV. Therefore, the adsorption free energy related to the HER mechanism can be considered as

graphic file with name ao1c06730_m003.jpg 3

For an ideal catalyst, the value of ΔGH* should be zero. Here, Eads is the hydrogen binding energy and can be defined as

graphic file with name ao1c06730_m004.jpg 4

where Emonolayer+nH and Emonolayer+(n–1)H are the total energy of the monolayer with n and n – 1 adsorbed hydrogen atoms, respectively, and EH2 is the total energy of H2 molecules in the gas phase.

Based on Nørskov’s approach,58 the catalytic descriptor ΔGH can be used to determine the hydrogen evolution exchange current density (i0) in the form of a volcano-shaped diagram. At pH = 0, i0 can be calculated from the following equations as

graphic file with name ao1c06730_m005.jpg 5
graphic file with name ao1c06730_m006.jpg 6

where k0 and kB are the rate constant and Boltzmann constant, respectively, and k0 is set to 1.

Results and Discussion

The in-plane optimized lattice constants of VGe2N4 and NbGe2N4 monolayers are 3.00 and 3.09 Å, respectively, consistent with the previous study.50 Both the monolayers exhibit hexagonal crystal structure with the space group Pm2 (No. 187), in which a metal layer is sandwiched between Ge and N layers in the vertical directions and stacked in the sequence of N–Ge–N–M–N–Ge–N as shown in Figure 1a,b, where M represents the heavy metal (V or Nb). It can also be viewed as an MN2 monolayer sandwiched between two Ge–N layers. Figure 1a shows the bond distances between interlayer Ge–N atoms, between M–N atoms (at the middle portion of the unit cell), and the intralayer Ge–N atoms (at both top and bottom portions), which are denoted by d1, d2, and d3, respectively. For the optimized VGe2N4 (NbGe2N4) monolayer, these values are 1.88 Å (1.87 Å), 2.06 Å (2.14 Å), and 1.85 Å (1.85 Å), respectively. The dynamical stability of the monolayers is confirmed by the positive phonon frequency in the calculated phonon band structures (Figure 1c,d).

Figure 1.

Figure 1

Optimized structure of 2 × 2 × 1 supercell of V/NbGeN4 monolayer: (a) side view and (b) top view (blue: Ge atoms, red: N atoms, green: V/Nb atoms). Phonon spectra of VGe2N4 and NbGe2N4 monolayers are shown in (c) and (d), respectively.

The band structures of both the monolayers within the PBE functional are shown in Figure S1 (Supporting Information). The hybrid HSE06 functional is employed to calculate the correct band gap and gain more insight into the electronic properties because the PBE functional shows a small band gap (∼0.008 eV for VGe2N4 and ∼0.001 eV for NbGe2N4). The calculated HSE06 band gaps of VGe2N4 and NbGe2N4 monolayers are 0.69 and 0.48 eV, respectively (Figure S2 of Supporting Information). For the VGe2N4 monolayer, the valence band maximum is situated at the Γ-point, whereas the conduction band minimum is located at the K-point, resulting in an indirect band gap. The band gaps in the spin-up and spin-down states are 0.69 and 2.55 eV, respectively (Figure S2a). On the other hand, for the NbGe2N4 monolayer, the spin-up and spin-down band gap values are 1.46 and 2.57 eV, receptively (Figure S2b).

The Hubbard onsite Coulomb potential U is used to fix the empirically over-delocalization issue of d electrons of V and Nb atoms. The U value is determined according to the HSE06 band structure. The calculated band gap at different U is matched with the HSE06 band gap, listed in Supporting Information (Table S1). Table S1 shows that the band gap of the VGe2N4 monolayer is found to be 0.67 eV at U = 4.5 eV (Figure 2a) close to the HSE06 band gap. Similarly, the band gap of the NbGe2N4 monolayer at U = 4 eV (Figure 2b) matches with the HSE06 band gap. Therefore, U = 4.5 eV (for VGe2N4) and 4 eV (for NbGe2N4) are used to further investigate the electronic and catalytic activities. The band structure and projected density of states (PDOS) of monolayers are shown in Figure 2. For the VGe2N4 monolayer, the spin-up states contribute to the conduction band minimum and are mainly dominated by V-3d states, whereas the spin-down state is away from the Fermi level. The valence band maximum is symmetrically contributed by both the spins and formed by the interaction between V-3d and N-2p states (Figure 2a), which is 0.67 eV lower than the conduction band minima. The scenario is different for the NbGe2N4 case as the valence and conduction bands near the Fermi level are contributed by spin-up and spin-down states, respectively, where a strong hybridization between Nb-4d and N-2p states is observed (Figure 2b), and are separated from each other by the gap of value 0.44 eV. In the deeper regime of the valence band, the N-2p orbitals weakly hybridized with Ge-4p and 3d/4d states of V/Nb are present.

Figure 2.

Figure 2

Calculated electronic band structure and project density of states of (a) VGe2N4 (U = 4.5) and (b) NbGe2N4 (U = 4). Spin-up and spin-down bands are represented by solid magenta lines and dotted purple lines, respectively. In the partial density of states, the contributions of V-3d (Nb-4d), Ge-4p, and N-2p orbitals are represented by green, blue, and red lines, respectively. The black line represents the total DOS.

Hydrogen Evolution Reaction Activity

The calculated Gibbs free energy for hydrogen adsorption at N sites of VGe2N4 and NbGe2N4 monolayers are −0.29 and 0.19 eV, respectively, far from the ideal value. The strong (weak) binding of a hydrogen atom on the basal planes of VGe2N4 (NbGe2N4) monolayers prohibits them from being an efficient HER electrocatalyst. Therefore, N and Ge vacancies are introduced in the basal plane. The ΔGH* values for N and Ge vacancies of the NbGe2N4 monolayer are −0.39 and 0.88 eV, respectively. The higher absolute binding energy value (|ΔEH|) indicates that the hydrogen evolution and adsorption are difficult in N and Ge vacancies, respectively. Later, doping of B and C atoms at N sites increases the binding energy compared to the pristine case and hinders the use of the NbGe2N4 monolayer as a promising electrocatalyst. In addition, O and S dopings at the N site show high positive ΔGH*. However, P doping at the N site causes the ΔGH* value to be 0.11 eV, which is much lower than the previous cases, and also results in increased exchange current. Figure 3 demonstrates a volcano plot to show the best active site of the NbGe2N4 monolayer for the HER reaction. P doping at the N site represented by H_N(P@N) offers the highest exchange current density among all possible doped cases. The detailed structural and electronic properties of the P-doped NbGe2N4 monolayer are given in the Supporting Information (Figure S3–S4). The same defect and doping engineering in the VGe2N4 monolayer do not enhance the HER activity significantly. The ΔGH* values for different doped cases are provided in the Supporting Information (Table S2).

Figure 3.

Figure 3

Volcano curve for the exchange current density i0 as a function of ΔGH* for the NbGe2N4 monolayer. H_X (X = N, Ge, C, B, O, P, S) represents the structures containing an H atom adsorbed above an X atom. NV and GeV represent N vacancy and Ge vacancy defect structures, respectively. Y@N (Y = P, S, C, O, B) represent the structures with Y atom doped at the N site. (For example, H_S(S@N) represents the structure containing an H atom placed above an S atom where the S atom is placed in the place of the N site of the NbGe2N4 monolayer).

The biaxial strains (−3 to +3%) are studied to shift the d-band center to possess the optimal rate for HER59 and observe the change in ΔGH*. The application of biaxial strain changes the strength of bonding between hydrogen and the substrate. As a result, ΔGH* value changes with respect to strain. In both materials, when we increase the strength of the tensile strain, the binding energy of H with monolayers increases and the interatomic distance between the atoms of the monolayer increases. As a result, the interatomic forces between them decrease. Hence, electrons of the monolayer are more prone to bond with H atoms and increase their strength. After analyzing the DOS, we have found the Fermi level to shift toward lower energy with the increase in the strength of the tensile strain, which is shown in Figure 6. This may be the reason for the change in H binding energy with respect to strain. The charge density, density of states, band structures, and band gaps of the two materials under strain are given in the Supporting Information (see Figures S5–S9 and Table S3). For the VGe2N4 monolayer, ΔGH* reduces with compressive strain (Figure 4a). At −3.0% strain, the ΔGH* value is −0.09 eV for the hydrogen coverage of 1/4, which indicates that the VGe2N4 monolayer shows better catalytic performance under compressive strain. The ΔGH* values are −0.17, −0.25, −0.43, −0.48, and −0.62 eV under −2, −1, 1, 2, and 3% strains, respectively. By gradually increasing the strains from −3 to +3%, the value of ΔGH* is driven toward a higher negative value, which implies the tight binding of the hydrogen atom with tensile strain. However, for the NbGe2N4 monolayer, the scenario is reversed. For 1/4 hydrogen coverage, the ΔGH* reaches zero with the increasing tensile strain (Figure 4b). At 2 and 3% strain, the ΔGH* values are determined to be 0.090 and 0.015 eV, respectively, which reveals that the NbGe2N4 monolayer can be a promising electrocatalyst toward HER under tensile strain. The ΔGH* values are 0.150, 0.220, 0.241, and 0.242 eV under 1, −1, −2, and −3% strain respectively. In both cases, applying tensile strain increases the bonding between H and the substrate and compressive strain weakens the bonding. Figure 4 shows that with the same range of biaxial strain, ΔGH* changes more for VGe2N4 than for NbGe2N4.

Figure 6.

Figure 6

Calculated density of states of (a) VGe2N4 and (b) NbGe2N4 monolayers with adsorbed hydrogen at different strains (shown in red color). The gray dashed vertical line represents the Fermi level.

Figure 4.

Figure 4

Gibbs free energy diagram of (a) VGe2N4 and (b) NbGe2N4 monolayers at different strains ranging from −3 to +3%. Here 0% strain indicates the pristine case.

The calculated charge density of the interface between the hydrogen atom and the monolayers is presented for the pristine sample (Figure 5) and density of states with adsorbed hydrogen at different compressive and tensile strains (Figure 6). Furthermore, Bader charge analysis60 is used to calculate the charge transfer. Figure 5 shows a significant charge loss from hydrogen towards nitrogen atom for both monolayers, which amounts to 0.48 e and 0.42 e for VGe2N4 and NbGe2N4 monolayers, respectively. In addition, the V and Nb atoms gain some charge (Figure 5a,b). The amount of charge transfer (ΔQ) for adsorbed hydrogen atoms on the monolayers at different strains is shown in Table 1. At −3% strain, the charge accumulates at V atoms in the VGe2N4 monolayer, which shows better HER activity (Figure S5). However, for the NbGe2N4 monolayer, at +3% strain, very few charges accumulate at Nb atoms. Thus, the accumulation or depletion of charge near the metal atoms plays an active role in HER performance. The variation of binding energies of hydrogen to the monolayers under strains is explained by the density of states (Figure 6). From compressive to tensile strains, the downward shift in the Fermi level for both cases is responsible for the change in HER activity under biaxial strain.

Figure 5.

Figure 5

Charge density difference plot for H adsorption on (a) VGe2N4 and (b) NbGe2N4 monolayers at 0% strain. (The isosurface value is 0.005 e/A3.) Yellow and cyan colors represent electron accumulation and depletion regions, respectively.

Table 1. Calculated Charge Transfer (ΔQ) for Adsorbed Hydrogen Atom on Both Monolayers at Different Strains Using Bader Charge Analysis.

strain (%) –3 –2 –1 0 1 2 3  
VGe2N4 0.478 0.459 0.512 0.481 0.472 0.442 0.427 ΔQ (e)
NbGe2N4 0.447 0.478 0.448 0.421 0.471 0.437 0.412

Although the accumulated charges on the metal atoms are very low, their existence is clearly shown in Figure 5. As compared to V, a larger charge is accumulated in Nb atoms, which signifies that inner metal atoms also play a role in hydrogen bonding. Thus, we obtain different binding energies for the different monolayers, although the outer-most atoms (Ge and N) of both the layers are the same. In addition, under strain, the charge accumulation near the metal atoms also changes, indicating the affect of metal atoms on HER (Figure S5).

Conclusions

In summary, we have systematically investigated the catalytic properties of VGe2N4 and NbGe2N4 monolayers toward the hydrogen evolution reaction and found that the Gibbs free energy ΔGH* is −0.29 and 0.19 eV, respectively, for the pristine VGe2N4 and NbGe2N4 monolayers. P doping at the N site in NbGe2N4 yields a higher Gibbs free energy (ΔGH* = 0.11 eV). The strain has a dramatic affect on HER performance on both VGe2N4 and NbGe2N4 monolayers. The values of ΔGH* of −0.09 eV for VGe2N4 at −3% strain and 0.015 eV for NbGe2N4 at +3% strain imply the potential applications of both the monolayers as efficient electrocatalysts for hydrogen production. In addition, compressive strains (−1 to −3%) positively affect the HER activity of the VGe2N4 monolayer, which weakens the bonding between hydrogen and the substrate, resulting in ΔGH* being close to zero. However, the opposite trend is found for the NbGe2N4 monolayer where the tensile strain (1–3%) increases the interaction between hydrogen and the monolayer, leading to a very low ΔGH*. Thus, our work provides a direction to experimentalists to not only design VGe2N4 and NbGe2N4 monolayers as electrocatalysts for HER activity but also to explore other members of the MA2Z4 family.

Acknowledgments

N.S. acknowledges the financial support from the Khalifa University of Science and Technology through the start-up grant FSU-2020-11.

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.1c06730.

  • PBE band structure of VGe2N4 and NbGe2N4 monolayers; HSE band structure of VGe2N4 and NbGe2N4 monolayers; the values of band gaps for VGe2N4 and NbGe2N4 monolayers for different U and HSE functionals; ΔGH* for N vacancy VGe2N4 surface doped by different atoms; charge density difference plots for VGe2N4 and NbGe2N4 monolayers at different strains; total density of states (DOS) of the VGe2N4 monolayer under biaxial strains; band structures of the VGe2N4 monolayer under biaxial strains; band structures of the NbGe2N4 monolayer under biaxial strains; total density of states (DOS) of the NbGe2N4 monolayer under biaxial strains; the values of band gaps for VGe2N4 and NbGe2N4 monolayers for different biaxial strains; structural, electronic, and catalytic properties of P-doped NbGe2N4; optimized structure of the P-doped NbGe2N4 monolayer; the charge density difference for H-adsorbed-doped NbGe2N4 monolayer; and projected density of states (PDOS) of P doping at N vacancy of the NbGe2N4 monolayer (PDF)

The authors declare no competing financial interest.

Supplementary Material

ao1c06730_si_001.pdf (1.9MB, pdf)

References

  1. Debe M. K. Electrocatalyst Approaches and Challenges for Automotive Fuel Cells. Nature 2012, 486, 43–51. 10.1038/nature11115. [DOI] [PubMed] [Google Scholar]
  2. Zheng Y.; Jiao Y.; Jaroniec M.; Qiao S. Z. Advancing the Electrochemistry of the Hydrogen- Evolution Reaction through Combining Experiment. Angew. Chem., Int. Ed. 2015, 54, 52–65. 10.1002/anie.201407031. [DOI] [PubMed] [Google Scholar]
  3. Vesborg P. C. K.; Seger B.; Chorkendorff I. Recent Development in Hydrogen Evolution Reaction Catalysts and Their Practical Implementation. J. Phys. Chem. Lett. 2015, 6, 951–957. 10.1021/acs.jpclett.5b00306. [DOI] [PubMed] [Google Scholar]
  4. Chen J.; Xia G.; Jiang P.; Yang Y.; Li R.; Shi R.; Su J.; Chen Q. Active and Durable Hydrogen Evolution Reaction Catalyst Derived from Pd-Doped Metal-Organic Frameworks. ACS Appl. Mater. Interfaces 2016, 8, 13378–13383. 10.1021/acsami.6b01266. [DOI] [PubMed] [Google Scholar]
  5. Grigoriev S. A.; Millet P.; Fateev V. N. Evaluation of Carbon-Supported Pt and Pd Nanoparticles for the Hydrogen Evolution Reaction in PEM Water Electrolysers. J. Power Sources 2008, 177, 281–285. 10.1016/j.jpowsour.2007.11.072. [DOI] [Google Scholar]
  6. Skúlason E.; Karlberg G. S.; Rossmeisl J.; Bligaard T.; Greeley J.; Jónsson H.; Nørskov J. K. Density Functional Theory Calculations for the Hydrogen Evolution Reaction in an Electrochemical Double Layer on the Pt(111) Electrode. Phys. Chem. Chem. Phys. 2007, 9, 3241–3250. 10.1039/b700099e. [DOI] [PubMed] [Google Scholar]
  7. Jiao Y.; Zheng Y.; Jaroniec M.; Qiao S. Z. Design of Electrocatalysts for Oxygen- and Hydrogen-Involving Energy Conversion Reactions. Chem. Soc. Rev. 2015, 44, 2060–2086. 10.1039/C4CS00470A. [DOI] [PubMed] [Google Scholar]
  8. Chen L. X.; Chen Z. W.; Jiang M.; Lu Z.; Gao C.; Cai G.; Singh C. V. Insights on the Dual Role of Two-Dimensional Materials as Catalysts and Supports for Energy and Environmental Catalysis. J. Mater. Chem. A 2021, 9, 2018–2042. 10.1039/D0TA08649E. [DOI] [Google Scholar]
  9. Siahrostami S.; Tsai C.; Karamad M.; Koitz R.; García-Melchor M.; Bajdich M.; Vojvodic A.; Abild-Pedersen F.; Nørskov J. K.; Studt F. Two-Dimensional Materials as Catalysts for Energy Conversion. Catal. Lett. 2016, 146, 1917–1921. 10.1007/S10562-016-1837-Z. [DOI] [Google Scholar]
  10. Deng D.; Novoselov K. S.; Fu Q.; Zheng N.; Tian Z.; Bao X. Catalysis with Two-Dimensional Materials and Their Heterostructures. Nat. Nanotechnol. 2016, 11, 218–230. 10.1038/nnano.2015.340. [DOI] [PubMed] [Google Scholar]
  11. Karmodak N.; Andreussi O. Catalytic Activity and Stability of Two-Dimensional Materials for the Hydrogen Evolution Reaction. ACS Energy Lett. 2020, 5, 885–891. 10.1021/ACSENERGYLETT.9B02689. [DOI] [Google Scholar]
  12. Pan H. Metal Dichalcogenides Monolayers: Novel Catalysts for Electrochemical Hydrogen Production. Sci. Rep. 2014, 4, 5348 10.1038/srep05348. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Shi W.; Li G.; Wang Z. Triggering Catalytic Active Sites for Hydrogen Evolution Reaction by Intrinsic Defects in Janus Monolayer MoSSe. J. Phys. Chem. C 2019, 123, 12261–12267. 10.1021/acs.jpcc.9b01485. [DOI] [Google Scholar]
  14. Nigar S.; Zhou Z.; Wang H.; Imtiaz M. Modulating the Electronic and Magnetic Properties of Graphene. RSC Adv. 2017, 7, 51546–51580. 10.1039/c7ra08917a. [DOI] [Google Scholar]
  15. Gao G.; Sun Q.; Du A. Activating Catalytic Inert Basal Plane of Molybdenum Disulfide to Optimize Hydrogen Evolution Activity via Defect Doping and Strain Engineering. J. Phys. Chem. C 2016, 120, 16761–16766. 10.1021/ACS.JPCC.6B04692. [DOI] [Google Scholar]
  16. Huang H.; Fan X.; Singh D. J.; Zheng W. Modulation of Hydrogen Evolution Catalytic Activity of Basal Plane in Monolayer Platinum and Palladium Dichalcogenides. ACS Omega 2018, 3, 10058–10065. 10.1021/ACSOMEGA.8B01414. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Ouyang Y.; Ling C.; Chen Q.; Wang Z.; Shi L.; Wang J. Activating Inert Basal Planes of MoS2 for Hydrogen Evolution Reaction through the Formation of Different Intrinsic Defects. Chem. Mater. 2016, 28, 4390–4396. 10.1021/ACS.CHEMMATER.6B01395. [DOI] [Google Scholar]
  18. Hu L.; Sun Y.; Gong S.-J.; Zong H.; Yu K.; Zhu Z. Experimental and Theoretical Investigation on MoS2/MXene Heterostructure as an Efficient Electrocatalyst for Hydrogen Evolution in Both Acidic and Alkaline Media. New J. Chem. 2020, 44, 7902–7911. 10.1039/D0NJ00956C. [DOI] [Google Scholar]
  19. Gan J.; Li F.; Tang Q. Vacancies-Engineered M2CO2 MXene as an Efficient Hydrogen Evolution Reaction Electrocatalyst. J. Phys. Chem. Lett. 2021, 12, 4805–4813. 10.1021/ACS.JPCLETT.1C00917. [DOI] [PubMed] [Google Scholar]
  20. Wang X.; Wang C.; Ci S.; Ma Y.; Liu T.; Gao L.; Qian P.; Ji C.; Su Y. Accelerating 2D MXene Catalyst Discovery for the Hydrogen Evolution Reaction by Computer-Driven Workflow and an Ensemble Learning Strategy. J. Mater. Chem. A 2020, 8, 23488–23497. 10.1039/D0TA06583H. [DOI] [Google Scholar]
  21. Yoon Y.; Tiwari A. P.; Lee M.; Choi M.; Song W.; Im J.; Zyung T.; Jung H.-K.; Lee S.S.; Jeon S.; An K.-S. Enhanced Electrocatalytic Activity by Chemical Nitridation of Two-Dimensional Titanium Carbide MXene for Hydrogen Evolution. J. Mater. Chem. A 2018, 6, 20869–20877. 10.1039/C8TA08197B. [DOI] [Google Scholar]
  22. Zhang X.; Shao B.; Sun Z.; Gao Z.; Qin Y.; Zhang C.; Cui F.; Yang X. Platinum Nanoparticle-Deposited Ti3C2Tx MXene for Hydrogen Evolution Reaction. Ind. Eng. Chem. Res. 2020, 59, 1822–1828. 10.1021/ACS.IECR.9B05046. [DOI] [Google Scholar]
  23. Xie G.; Zhang K.; Guo B.; Liu Q.; Fang L.; Gong J. R. Graphene-Based Materials for Hydrogen Generation from Light-Driven Water Splitting. Adv. Mater. 2013, 25, 3820–3839. 10.1002/ADMA.201301207. [DOI] [PubMed] [Google Scholar]
  24. Nemiwal M.; Zhang T. C.; Kumar D. Graphene-Based Electrocatalysts: Hydrogen Evolution Reactions and Overall Water Splitting. Int. J. Hydrogen Energy 2021, 46, 21401–21418. 10.1016/J.IJHYDENE.2021.04.008. [DOI] [Google Scholar]
  25. Yan H.; Xie Y.; Jiao Y.; Wu A.; Tian C.; Zhang X.; Wang L.; Fu H. Holey Reduced Graphene Oxide Coupled with an Mo2N–Mo2C Heterojunction for Efficient Hydrogen Evolution. Adv. Mater. 2018, 30, 1704156 10.1002/ADMA.201704156. [DOI] [PubMed] [Google Scholar]
  26. Li J.-S.; Wang Y.; Liu C.-H.; Li S.-L.; Wang Y.-G.; Dong L.-Z.; Dai Z.-H.; Li Y.-F.; Lan Y.-Q. Coupled Molybdenum Carbide and Reduced Graphene Oxide Electrocatalysts for Efficient Hydrogen Evolution. Nat. Commun. 2016, 7, 11204 10.1038/ncomms11204. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Guan M.; Wang C.; Li S.; Du H.; Yuan Y. Understanding the Enhanced Electrocatalytic Hydrogen Evolution via Integrating Electrochemically Inactive G-C3N4: The Effect of Interfacial Engineering. ACS Sustainable Chem. Eng. 2020, 8, 10313–10320. 10.1021/ACSSUSCHEMENG.0C03938. [DOI] [Google Scholar]
  28. Chen Y.; Yu G.; Chen W.; Liu Y.; Li G.-D.; Zhu P.; Tao Q.; Li Q.; Liu J.; Shen X.; et al. Highly Active, Nonprecious Electrocatalyst Comprising Borophene Subunits for the Hydrogen Evolution Reaction. J. Am. Chem. Soc. 2017, 139, 12370–12373. 10.1021/JACS.7B06337. [DOI] [PubMed] [Google Scholar]
  29. Lu J.; Zhang X.; Liu D.; Yang N.; Huang H.; Jin S.; Wang J.; Chu P. K.; Yu X.-F. Modulation of Phosphorene for Optimal Hydrogen Evolution Reaction. ACS Appl. Mater. Interfaces 2019, 11, 37787–37795. 10.1021/ACSAMI.9B13666. [DOI] [PubMed] [Google Scholar]
  30. Lou H.; Yu T.; Ma J.; Zhang S.; Bergara A.; Yang G. Achieving High Hydrogen Evolution Reaction Activity of a Mo2C Monolayer. Phys. Chem. Chem. Phys. 2020, 22, 26189–26199. 10.1039/D0CP05053A. [DOI] [PubMed] [Google Scholar]
  31. Zhao X.; Yang X.; Singh D.; Panda P. K.; Luo W.; Li Y.; Ahuja R. Strain-Engineered Metal-Free h-B2O Monolayer as a Mechanocatalyst for Photocatalysis and Improved Hydrogen Evolution Reaction. J. Phys. Chem. C 2020, 124, 7884–7892. 10.1021/ACS.JPCC.0C00834. [DOI] [Google Scholar]
  32. Gan Y.; Xue X.-X.; Jiang X.-X.; Xu Z.; Chen K.; Yu J.-F.; Feng Y. Chemically Modified Phosphorene as Efficient Catalyst for Hydrogen Evolution Reaction. J. Phys.: Condens. Matter 2019, 32, 025202 10.1088/1361-648X/AB482B. [DOI] [PubMed] [Google Scholar]
  33. Vineesh T. V.; Yarmiayev V.; Zitoun D. Tailoring the Electrochemical Hydrogen Evolution Activity of Cu3P through Oxophilic Surface Modification. Electrochem. Commun. 2020, 113, 106691 10.1016/J.ELECOM.2020.106691. [DOI] [Google Scholar]
  34. Yilmaz G.; Yang T.; Du Y.; Yu X.; Feng Y. P.; Shen L.; Ho G. W. Stimulated Electrocatalytic Hydrogen Evolution Activity of MOF-Derived MoS2 Basal Domains via Charge Injection through Surface Functionalization and Heteroatom Doping. Adv. Sci. 2019, 6, 1900140 10.1002/ADVS.201900140. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Wang M.; Song R.; Zhang X.; Liu G.; Xu S.; Xu Z.; Liu J.; Qiao G. Defects Engineering Promotes the Electrochemical Hydrogen Evolution Reaction Property of Phosphorene Surface. Int. J. Hydrogen Energy 2021, 46, 1913–1922. 10.1016/J.IJHYDENE.2020.10.039. [DOI] [Google Scholar]
  36. Ye G.; Gong Y.; Lin J.; Li B.; He Y.; Pantelides S. T.; Zhou W.; Vajtai R.; Ajayan P. M. Defects Engineered Monolayer MoS2 for Improved Hydrogen Evolution Reaction. Nano Lett. 2016, 16, 1097–1103. 10.1021/ACS.NANOLETT.5B04331. [DOI] [PubMed] [Google Scholar]
  37. Sathe B. R.; Zou X.; Asefa T. Metal-Free B-Doped Graphene with Efficient Electrocatalytic Activity for Hydrogen Evolution Reaction. Catal. Sci. Technol. 2014, 4, 2023–2030. 10.1039/C4CY00075G. [DOI] [Google Scholar]
  38. Shinde S. S.; Sami A.; Lee J.-H. Nitrogen- and Phosphorus-Doped Nanoporous Graphene/Graphitic Carbon Nitride Hybrids as Efficient Electrocatalysts for Hydrogen Evolution. ChemCatChem 2015, 7, 3873–3880. 10.1002/CCTC.201500701. [DOI] [Google Scholar]
  39. Zhang P.; Xu X.; Song E.; Hou X.; Yang X.; Mi J.; Huang J.; Stampfl C. Transition Metal-Doped α-Borophene as Potential Oxygen and Hydrogen Evolution Electrocatalyst: A Density Functional Theory Study. Catal. Commun. 2020, 144, 106090 10.1016/J.CATCOM.2020.106090. [DOI] [Google Scholar]
  40. Chen X.; Wang G. Tuning the Hydrogen Evolution Activity of MS2 (M = Mo or Nb) Monolayers by Strain Engineering. Phys. Chem. Chem. Phys. 2016, 18, 9388–9395. 10.1039/C5CP06475A. [DOI] [PubMed] [Google Scholar]
  41. Hong Y.-L.; Liu Z.; Wang L.; Zhou T.; Ma W.; Xu C.; Feng S.; Chen L.; Chen M.-L.; Sun D.-M.; et al. Chemical Vapor Deposition of Layered Two-Dimensional MoSi2N4 Materials. Science 2020, 369, 670–674. 10.1126/SCIENCE.ABB7023. [DOI] [PubMed] [Google Scholar]
  42. Yu J.; Zhou J.; Wan X.; Li Q. High Intrinsic Lattice Thermal Conductivity in Monolayer MoSi2N4. New J. Phys. 2021, 23, 033005 10.1088/1367-2630/ABE8F7. [DOI] [Google Scholar]
  43. Yao H.; Zhang C.; Wang Q.; Li J.; Yu Y.; Xu F.; Wang B.; Wei Y. Novel Two-Dimensional Layered MoSi2Z4 (Z = P, As): New Promising Optoelectronic Materials. Nanomaterials 2021, 11, 559 10.3390/NANO11030559. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Jian C.; Ma X.; Zhang J.; Yong X. Strained MoSi2N4 Monolayers with Excellent Solar Energy Absorption and Carrier Transport Properties. J. Phys. Chem. C 2021, 125, 15185–15193. 10.1021/acs.jpcc.1c03585. [DOI] [Google Scholar]
  45. Ray A.; Tyagi S.; Singh N.; Schwingenschlögl U. Inducing Half-Metallicity in Monolayer MoSi2N4. ACS Omega 2021, 6, 30371–30375. 10.1021/ACSOMEGA.1C03444. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Yadav A.; Kangsabanik J.; Singh N.; Alam A. Novel Two-Dimensional MA2N4Materials for Photovoltaic and Spintronic Applications. J. Phys. Chem. Lett. 2021, 12, 10120–10127. 10.1021/acs.jpclett.1c02650. [DOI] [PubMed] [Google Scholar]
  47. Liu Y.; Ji Y.; Li Y. Multilevel Theoretical Screening of Novel Two-Dimensional MA2Z4Family for Hydrogen Evolution. J. Phys. Chem. Lett. 2021, 12, 9149–9154. 10.1021/acs.jpclett.1c02487. [DOI] [PubMed] [Google Scholar]
  48. Zang Y.; Wu Q.; Du W.; Dai Y.; Huang B.; Ma Y. Activating Electrocatalytic Hydrogen Evolution Performance of Two-Dimensional MSi2N4 (M = Mo,W): A Theoretical Prediction. Phys. Rev. Mater. 2021, 5, 045801 10.1103/PhysRevMaterials.5.045801. [DOI] [Google Scholar]
  49. Xiao C.; Sa R.; Cui Z.; Gao S.; Du W.; Sun X.; Zhang X.; Li Q.; Ma Z. Enhancing the Hydrogen Evolution Reaction by Non-Precious Transition Metal (Non-Metal) Atom Doping in Defective MoSi2N4 Monolayer. Appl. Surf. Sci. 2021, 563, 150388 10.1016/J.APSUSC.2021.150388. [DOI] [Google Scholar]
  50. Wang L.; Shi Y.; Liu M.; Zhang A.; Hong Y.-L.; Li R.; Gao Q.; Chen M.; Ren W.; Cheng H.-M.; et al. Intercalated Architecture of MA2Z4 Family Layered van Der Waals Materials with Emerging Topological, Magnetic and Superconducting Properties. Nat. Commun. 2021, 12, 2361 10.1038/s41467-021-22324-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Nørskov J. K.; Bligaard T.; Logadottir A.; Kitchin J. R.; Chen J. G.; Pandelov S.; Stimming U. Trends in the Exchange Current for Hydrogen Evolution. J. Electrochem. Soc. 2005, 152, J23. 10.1149/1.1856988. [DOI] [Google Scholar]
  52. Kresse G.; Hafner J. Norm-Conserving and Ultrasoft Pseudopotentials for First-Row and Transition Elements. J. Phys.: Condens. Matter 1994, 6, 8245–8257. 10.1088/0953-8984/6/40/015. [DOI] [Google Scholar]
  53. Hafner J. Ab-Initio Simulations of Materials Using VASP: Density-Functional Theory and Beyond. J. Comput. Chem. 2008, 29, 2044–2078. 10.1002/jcc.21057. [DOI] [PubMed] [Google Scholar]
  54. Perdew J. P.; Burke K.; Ernzerhof M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865. 10.1103/PhysRevLett.77.3865. [DOI] [PubMed] [Google Scholar]
  55. Heyd J.; Scuseria G. E.; Ernzerhof M. Hybrid Functionals Based on a Screened Coulomb Potential. J. Chem. Phys. 2003, 118, 8207–8215. 10.1063/1.1564060. [DOI] [Google Scholar]
  56. Togo A.; Tanaka I. First Principles Phonon Calculations in Materials Science. Scr. Mater. 2015, 108, 1–5. 10.1016/J.SCRIPTAMAT.2015.07.021. [DOI] [Google Scholar]
  57. Tsai C.; Abild-Pedersen F.; Nørskov J. K. Tuning the MoS2 Edge-Site Activity for Hydrogen Evolution via Support Interactions. Nano Lett. 2014, 14, 1381–1387. 10.1021/NL404444K. [DOI] [PubMed] [Google Scholar]
  58. Hammer B.; Norskov J. K. Why Gold Is the Noblest of All the Metals. Nature 1995, 376, 238–240. 10.1038/376238a0. [DOI] [Google Scholar]
  59. You B.; Tang M. T.; Tsai C.; Abild-Pedersen F.; Zheng X.; Li H. Enhancing Electrocatalytic Water Splitting by Strain Engineering. Adv. Mater. 2019, 31, 1807001 10.1002/ADMA.201807001. [DOI] [PubMed] [Google Scholar]
  60. Henkelman G.; Arnaldsson A.; Jónsson H. A Fast and Robust Algorithm for Bader Decomposition of Charge Density. Comput. Mater. Sci. 2006, 36, 354–360. 10.1016/j.commatsci.2005.04.010. [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

ao1c06730_si_001.pdf (1.9MB, pdf)

Articles from ACS Omega are provided here courtesy of American Chemical Society

RESOURCES