Defect engineering in the MA2Z4 monolayer family for enhancing the hydrogen evolution reaction: first-principles calculations†
Received
7th October 2022
, Accepted 12th November 2022
First published on 14th November 2022
Abstract
Water electrolysis is a sustainable and clean method to produce hydrogen fuel via the hydrogen evolution reaction (HER). Using effective and low-cost electrocatalysts for the HER to substitute expensive noble metals is highly desired. In this study, by using first-principles calculations, we designed a defective MA2Z4 monolayer family as two-dimensional (2D) electrocatalysts for the HER, and their stability, electronic properties, and catalytic performance were investigated. As the most representative material of the MA2Z4 monolayer family, MoSi2N4 was first successfully synthesized in the experiment [Y. L. Hong et al., Science, 2020, 369, 670–674]. Our results reveal that vacancy regulation at the outer Z atom (VZ) can obviously enhance the catalytic activity toward the HER, compared with the pristine MA2Z4. It is also shown that ΔGH* can be close to 0 eV with VP vacancies at active site 1, which should be the optimal performance for a HER catalyst. Moreover, we demonstrated that the HER performance prefers the Volmer–Heyrovsky mechanism. Our study provides a strategy for designing MA2Z4 monolayer family electrocatalysts, which are predicted to be employed in HER catalysts with low cost and high performance.
1. Introduction
Hydrogen has become one of the most promising alternatives to substitute fossil fuels, owing to its zero pollution, high combustion value, high energy density, and renewable properties.1–4 There are many approaches to produce hydrogen, such as the decomposition of coal, water, and hydrocarbon. Water electrolysis is a sustainable, clean, and low-cost method to produce H2, and electrocatalysts can enhance the efficiency of water splitting observably.5 For the hydrogen evolution reaction (HER), platinum and its alloys are considered potential catalysts to activate the HER owing to the small Tafel slope, low overpotential, and high exchange current density.6–8 However, the high price and scarcity hamper the exploration of nonnoble metal-based catalysts having HER catalytic activity comparable to it.
In recent years, lots of nonnoble metal-based catalysts, such as carbides,9–12 borides,13,14 phosphides,15,16 sulfides,17–19 nitrides20–22 and so on, have been discovered and made significant progress as a potential catalyst for activating the HER. Among these materials, two-dimensional (2D) nano materials provide new opportunities for the HER because of the compelling structural and electronic properties. Graphene-based materials and transition metal dichalcogenides (TMDs) have gained much attention as 2D electrocatalysts for the HER due to their high electronic conductivity and uniformly distributed active sites.23–25 Recently, a new type of 2D material MA2Z4 monolayer family,26–28 in which M represents Mo, W or V, A represents Si or Ge, and Z represents an early transition stands for N, P, or As, have been proposed and exhibited high strength and remarkable stability. Among these materials, MoSi2N4 and WSi2N4 were first successfully synthesized by chemical vapor deposition with high strength (66 GPa), semiconductor properties (band gap 1.94 eV) and excellent stability.29,30 The unique and adjustable structure of septuple Z–A–Z–M–Z–A–Z (7 atomic layer thicknesses) possesses great possibilities of integrating various properties. Because of the fascinating properties like other 2D materials, such as good ambient stability, high carrier mobility, and broad hole and electron mobilities,31 it is reasonably speculated that highly efficient HER catalysts may exist in the MA2Z4 monolayer family. Therefore, searching ideal electrocatalysts from the family by high-throughput screening must have great significance for the development of HER catalysts, but limited efforts have been made in this field.
In this study, using first-principles calculation, we designed and demonstrated defective MA2Z4, investigated their stability and electronic properties, and evaluated their performance as HER electrocatalysts. Our results reveal that vacancy regulation at the outer Z atom (VZ) can obviously enhance the catalytic activity toward the HER, compared with pristine MA2Z4. It is also shown that ΔGH* can be close to 0 eV with VP vacancies at active site 1, which should be the optimal performance for a HER catalyst, so VP vacancies MA2P4 should have optimal ΔGH* and activation energy barrier for the rate-determining step among the family, and it exhibits more favorable performance. We also demonstrate that the Volmer–Heyrovsky mechanism is more preferred for the HER on defective MA2Z4. We compared our results with that of other researchers on MA2Z4, and it can be found that the VZ defect is more effective for MA2Z4 in the catalysis of the HER. Thus, our effort on defective MA2Z4 makes it a highly promising electrocatalyst for th eHER, and our findings provide a deep understanding in designing efficient and durable electrocatalysts.
2. Computational methods
Our first-principles calculations were performed using the Vienna Ab initio Simulation Package (VASP).32 The projected augmented wave (PAW) potentials were used to analyze the interactions between core electrons and valence electrons.33 The electron exchange–correlation interactions were described using the Perdew–Burke–Ernzerhof (PBE) functional within generalized gradient approximation (GGA).34 The DFT-D3 exchange–correlation functional was introduced in structural optimization to take the van der Waals interaction into account. The vacuum space along the z-direction was set to 20 Å in order to eliminate the interactions between MA2Z4 and its periodic images.
The plane-wave energy cutoff was set to be 500 eV. The convergence criterion was set as 10−5 eV for total energy. All the atomic positions and lattice structures were fully relaxed with the threshold of a maximum force of 0.02 eV Å−1. In order to ensure the accuracy and efficiency of the calculation, a Gamma-centered k-point mesh with a Monkhorst–Pack method 5 × 5 × 1 was employed for all considered structures after the convergence test.35 The amount of charge transfer between the C atoms and H atoms was calculated using Bader code.36 We also calculated H* adsorption energy barriers using the climbing image-nudged elastic band (CI-NEB) method.37 The CI-NEB is an efficient method to determine the minimum energy path and saddle points between a given initial and final position,38,39 and in our CI-NEB calculations, the initial and the final structures were fully optimized.
The adsorption energy (ΔEH) is defined as
| | (1) |
where
E(*H) and
E(*) are the total energy of structures with and without hydrogen adsorption, respectively, and
E(H
2) is the total energy of a H
2 molecule.
The Gibbs free energy (ΔGH*) is defined as:
| ΔGH* = ΔEH + ΔEZPE − TΔSH | (2) |
where Δ
EH is the adsorption energy, Δ
EZPE is the difference in zero-point energy,
T is the temperature (298.15 K) and Δ
SH is the entropy difference of adsorbed H and H in the gas phase. We approximated the entropy of hydrogen adsorption as
, where
is the entropy of gas phase H
2 under standard conditions, and
TΔ
SH was set to be −0.202 eV (ref.
40 and
41) after calculation in this study.
3. Results and discussion
3.1 Geometries and stability of MA2Z4
The optimized structures of MA2Z4 (M = Mo, W, V; A = Si, Ge; Z = N, P, As) are shown in Fig. 1(a). The MA2Z4 monolayer is a seven-layer honeycomb structure, in which the order of the M, A, and Z atom covalent bond is Z–A–Z–M–Z–A–Z, and the layers are connected by van der Waals (vdW) forces.42 In the pristine structure, the adsorption sites can be divided into four types. The site 1 is the top site of Z atoms, the site 2 is the top site of inside Z atoms, site 3 is the top site of A atoms, and the site 4 is the top site of M atoms.
|
| Fig. 1 Top and side views of (a) optimized pristine MA2Z4, (b) VM-MA2Z4, (c) VA-MA2Z4, and (d) VZ-MA2Z4. Note that the dotted red circle represents the defect site and numbers represent possible H adsorption sites. Purple, blue, and pink balls represent M, A, and Z atoms, respectively. The red diamond represents the primitive structure, and t is the thickness of MA2Z4 (for interpretation of the references to color in this figure legend, the reader is referred to the web version of this article). | |
For pristine MA2Z4, the calculated lattice constants, and the thickness of the monolayer layer are listed in Table 1 (while A–Z (d1, d2) and M–Z (d3) bond lengths; A–Z–A (θ1), Z–A–Z (θ2) and Z–M–Z (θ3, θ4) angles are listed in Table S1†). Clearly, the lattice constants exhibit an increasing trend with the increasing atomic radius of either M, A, or Z. For instance, for W-based MA2Z4, the lattice constant (a = b) has the order of WSi2N4 (2.54 Å) < WGe2N4 (3.04 Å) < WSi2P4 (3.46 Å) < WGe2P4 (3.55 Å) < WSi2As4 (3.60 Å) < WGe2As4 (3.70 Å). The thicknesses (t) of MoSi2N4 (7.00 Å), MoSi2P4 (9.34 Å), MoSi2As4 (9.90 Å), and WSi2N4 (7.01 Å), and the lattice constants (a = b) of MoSi2N4 (2.90 Å), MoSi2P4 (3.45 Å), MoSi2As4 (3.58 Å), and WSi2N4 (2.54 Å) agree well with previous reports.43–47 According to the symmetry of MA2Z4 structures, we consider three kinds of vacancies, which are outside the Z atom vacancy, A atom vacancy, and M atom vacancy to verify the performance of defective structures in Fig. 1(b)–(d). The dotted red circle represents the M atomic vacancy, A atomic vacancy, and outside Z atomic vacancies, respectively. We select four atoms next to the vacancy as the H atom adsorption sites; the specific positions are marked with numbers in Fig. 1(b)–(d).
Table 1 The optimized geometric and electronic structure parameters of the MA2Z4 monolayer including the lattice constant (a, b); the thickness of monolayer layer (t) and the corresponding data in previous reports
Structure |
a = b (Å) |
Reported a = b (Å) |
t (Å) |
Reported t (Å) |
Structure |
a = b (Å) |
Reported a = b (Å) |
t (Å) |
Reported t (Å) |
MoSi2N4 |
2.90 |
2.89 (ref. 44) |
7.00 |
7.01 (ref. 45) |
WGe2N4 |
3.04 |
3.09 (ref. 44) |
7.48 |
— |
MoSi2P4 |
3.45 |
3.46 (ref. 48) |
9.34 |
9.37 (ref. 46) |
WGe2P4 |
3.55 |
3.54 (ref. 48) |
23.10 |
— |
MoSi2As4 |
3.58 |
3.61 (ref. 48) |
9.90 |
9.94 (ref. 46) |
WGe2As4 |
3.70 |
3.69 (ref. 48) |
10.20 |
— |
MoGe2N4 |
3.03 |
3.02 (ref. 48) |
7.48 |
7.39 (ref. 46) |
VSi2N4 |
2.86 |
2.88 (ref. 28) |
6.87 |
— |
MoGe2P4 |
3.53 |
3.53 (ref. 48) |
9.63 |
— |
VSi2P4 |
3.48 |
3.48 (ref. 48) |
9.15 |
9.25 (ref. 49) |
MoGe2As4 |
3.67 |
3.69 (ref. 48) |
10.20 |
— |
VSi2As4 |
3.58 |
3.65 (ref. 44) |
9.78 |
— |
WSi2N4 |
2.90 |
2.91 (ref. 47) |
7.01 |
7.01 (ref. 47) |
VGe2N4 |
3.00 |
3.01 (ref. 28) |
7.36 |
— |
WSi2P4 |
3.46 |
3.46 (ref. 48) |
9.32 |
— |
VGe2P4 |
3.55 |
3.56 (ref. 48) |
9.48 |
— |
WSi2As4 |
3.60 |
3.61 (ref. 48) |
9.91 |
— |
VGe2As4 |
3.66 |
3.72 (ref. 48) |
10.40 |
— |
Then, we calculate the formation energy to verify the stability of the structure. The results are shown in Fig. 2. The formation energy of vacancy is calculated using the following formula:
| Ef = Evacancy − Eprimitive + Ev-atom | (3) |
where
Evacancy is the energy of MA
2Z
4 with a vacancy, and
Eprimitive is the energy of primitive MA
2Z
4.
Ev-atom is the chemical potential of the vacancy atom, which can be obtained from the corresponding elementary unit cell or gas molecule.
45 As we know, a lower formation energy indicates an energetically favourable and feasible defective structure, so we try to find the more stable structure among three types of vacancies of all defective structures. Our calculation results show that the formation energy of V
Si-MoSi
2N
4 is 0.33 eV per atom, which agrees well with previous reports (0.32 eV per atom),
45 and the defect formation energies of MoGe
2P
4, WGe
2P
4 and VGe
2P
4 are the three smallest energies among all 18 structures, which are also smaller than the previous reported transition metal doped MoSi
2N
4 (0.29 eV per atom)
45 and non-transitional metal MoSi
2N
4 (0.34 eV per atom),
45 showing that it is more favourable in experimental preparation. We will therefore focus on MA
2P
4 in the following investigation.
|
| Fig. 2 The formation energies of MA2Z4 (M = Mo, W, V; A = Si, Ge; Z = N, P, As). Note that the orange histograms represent the formation energy of VM, cyan histograms represent the formation energy of VA, and the yellow histograms represent the formation energy of VZ-MA2Z4, respectively. | |
3.2 Activity of defect MA2Z4 toward the HER
To evaluate the catalytic activity of the vacancy defects of MA2Z4, we calculated the ΔGH* of vacancies at M, A, inside Z atoms, and outer Z atoms, respectively. ΔGH* is an important descriptor to evaluate the performance of the catalyst; an excellent catalyst should keep ΔGH* near zero, marked in dark purple in Fig. 3. We can see that the adsorption performance of MA2Z4 materials is improved after vacancy regulation. We also find that the vacancy at the M atom (VM) makes some structure distortion, such as VMo-MoSi2As4 and VV-VGe2X4,;the vacancy at the A atom (VA) makes the ΔGH* value more negative, which indicates that it is more difficult for hydrogen atoms to be desorbed. So, we demonstrate that MA2Z4 materials are not suitable for vacancy regulation at the M or A atom. The results show that vacancy regulation at the outer Z atom (VZ) can enhance the adsorption of the H atom, and all structures are stable after adsorption, which are consistent with the previously calculated formation energy. The results also show that the catalytic performance of the original MA2P4 and VP-MA2P4 are better than that of other materials, which means that the vacancy defects at the outer P atom are the most suitable regulation method for MA2Z4 materials.
|
| Fig. 3 Calculations of the adsorption performance of MA2Z4 under origin MA2Z4 (a) and under the N vacancy (b); P vacancy (c); As vacancy (d) regulation and the Gibbs free energy of the H atom at sites 1, 2, 3 and 4. Catalytic performance can be expressed by different colors, in which darker purple indicates stronger catalytic activity and white indicates weak catalytic activity. | |
We then studied the differences of catalytic performance between different active sites after the best vacancy regulation methods are selected. As mentioned before, using the MA2P4 structure as an example, we find that the catalytic activity of all sites is improved after VP was introduced, and the site 1 (VP1) showed the best catalytic activity, as shown Fig. 3.
According to the calculation results, we selected MoGe2P4, WGe2P4 and VGe2P4 structures to calculate their corresponding volcanic curves to observe the reaction rate, as shown in Fig. 4.41 We can see that both WGe2P4 and VGe2P4 are closer to Pt at the volcano summit, especially in the case of VP, reflecting their strong hydrogen evolution ability. It can be seen from Fig. 4(d) that ΔGH* can be closer to 0 than primitive structures. We can see that the ΔGH* of WGe2P4 at 1 site is 0.95 eV, and the ΔGH* of WGe2P4 becomes 0.05 eV after VP was introduced; the ΔGH* of MoGe2P4 is 0.91 eV, and the ΔGH* of MoGe2P4 becomes −0.02 eV after VP was introduced while ΔGH* of VGe2P4 decreased by 0.1 eV (from 0.13 eV to 0.03 eV).
|
| Fig. 4 The volcanic curve of (a) MoGe2P4, (b) WGe2P4, and (c) VGe2P4 for each vacancy. (d) Gibbs free energy versus the reaction coordinate of the HER for MoGe2P4, WGe2P4, and VGe2P4 when VP was introduced on site 1. | |
The favorable and unfavorable factors for HER performance are listed as follows in Table 2. We investigated the properties of the pristine structure firstly, which is found to be inert for the HER, and then we investigated the properties of the defective material through the introduction of vacancies, and the results show that MoGe2P4 with a P vacancy and outside P as the H adsorption site is the best structure among them. The calculation results reveal that ΔGH* decreases significantly, demonstrating that VP-MA2Z4 is very effective in reducing ΔGH*; this is consistent with our previous discussion. Thus, our results clearly suggest that the ΔGH* of VP-MA2Z4 can be manipulated to achieve the optimal HER activity. In addition, the ΔGH* of other materials in the catalytic active region after vacancy introduction are presented in Fig. S1 of the ESI.†
Table 2 The vacancy type, ΔGH*, and H adsorption site of MoGe2P4, WGe2P4 and VGe2P4 are shown in the following table
Structure |
Vacancy type |
ΔGH* |
Adsorption site |
MoGe2P4 |
Pristine |
0.74 |
2 |
M |
0.09 |
3 |
A |
−1.31 |
2 |
Z |
−0.02 |
1 |
WGe2P4 |
Pristine |
0.43 |
2 |
M |
−0.08 |
3 |
A |
0.32 |
1 |
Z |
0.05 |
1 |
VGe2P4 |
Pristine |
0.01 |
4 |
M |
— |
— |
A |
−0.95 |
2 |
Z |
0.03 |
1 |
3.3 Origin of the HER catalytic activity
3.3.1 DOS and band structures.
In order to understand the reasons for improving HER performance, we study the electronic properties of VP vacancy structures, including their band structure, projection state density, differential charge density and Bader charge analysis. For the VA2Z4 series, they all show metallic properties; it can be found that when a vacancy is introduced, impurity states near the Fermi level increase, thus improving H* adsorption strength.20–22 For the MoA2Z4 and WA2Z4 series, they present semiconductor properties. Among them, MoSi2N4, MoGe2N4, MoGe2P4, MoGe2As4, WSi2N4 and WGe2N4 are indirect band gap semiconductors, while MoSi2P4, MoSi2As4, WSi2P4, WSi2As4, WGe2P4 and WGe2As4 are direct band gap semiconductors. For the MA2N4 series, the band gaps are relatively larger (>1.20 eV), and the band gaps are smaller for MGe2P4 and MGe2As4 (<0.50 eV). When the vacancy occurs, the impurity states of several catalysts begin to aggregate near the Fermi level, and unmatched electrons may appear on the plane, and these unstable electrons will increase the conductivity of the catalyst, similarly reducing the band gap of the catalyst, and improving the conductivity of the catalysts and the activity of hydrogen evolution. This is consistent with the result that the value of ΔGH* begins to approach zero.
We study the partial density of state (PDOS) MGe2P4 (M = Mo, W, V) with VP vacancies and corresponding pristine structures, as shown in Fig. 5. The curves we describe in the PDOS diagram are all from atoms near vacancies. It can be seen from the figures that compared with the pristine structure, the VP vacancy structures show larger occupied states near the Fermi level, which are mainly derived from M-d orbitals, and the occupied states near the Fermi level gradually increase from MoGe2P4 to WGe2P4 and then to VGe2P4. When M = W and V, the P-p orbitals have a significant elevation near the Fermi level, which is related to the absence of the P atom in the outer layer. Meanwhile, we can see that when M = W, the Ge-p orbital also gets some lift; this indicates that the introduction of vacancies can effectively improve the dominance of d electrons and p electrons near the Fermi level, thus improving the catalytic activity.
|
| Fig. 5 The band and PDOS diagram of (a, c and e) pristine MGe2P4 (M = Mo, W, V) and the (b, d and f) MGe2P4 (M = Mo, W, V) with the VP vacancy. | |
3.3.2 Electron density difference and charge transfer.
The Bader charge of H* and differential charge density are calculated for exploring the charge transfer mechanism between H* and defective structures and the relationship with hydrogen evolution activity. We calculated the Bader charge of adsorbed H* on defective MA2Z4 with VZ vacancies, as shown in Fig. 6, the M atoms of Mo, W, and V are represented in green, blue, and red, respectively. We find that electrons are transferred from H* to the N atom, while the electrons are transferred from the P or As atom to H*, because of the difference of the electronegativity of the Z atoms. The electronegativity of the N atom is greater than that of P and As, which affects the charge transfer when H* is adsorbed, resulting in the value of ΔGH*. We further show that there is a larger charge transfer between H* and Z atoms when VZ vacancies are introduced, which means that the stabilization of the H* species in HER performance may originate from the enhanced charge density. As mentioned above, the ΔGH* calculations also suggested that VZ could much more efficiently enhance the HER activity, especially VP structures.
|
| Fig. 6 The correlation of ΔGH* and Bader charge analysis of adsorbed H in defective structures. | |
To further explore the charge transfer mechanism between defective structures and H*, the differential charge transfer density (Δρ(r)) is calculated. Δρ(r) is calculated as
| Δρ(r) = ρcat+H(r) − ρcat(r) − ρH(r) | (4) |
where
ρcat+H(
r),
ρcat(
r), and
ρH(
r) denote charge density of the catalyst with adsorbed H, without H and H atoms, respectively. As mentioned above, we calculated the chosen V
P structure, and the calculated Δ
ρ(r) results of MoGe
2P
4, WGe
2P
4, and VSi
2P
4 are shown in
Fig. 7. It is shown that the electrons accumulate around H atoms and reduce around the P atoms which are bonded to H atoms, indicating a charge transfer from MA
2P
4 to H*, which is consistent with the Bader charge analysis.
|
| Fig. 7 Differential charge density of (a) MoGe2P4, (b) WGe2P4, and (c) VGe2P4. Note that the yellow and blue colors represent charge accumulation and depletion respectively. The isosurface level is 0.002 e Bohr−3. | |
3.4 The reaction pathways of defective MA2Z4
The hydrogen evolution reaction is a multi-step electrochemical reaction, which is divided into two cases: Volmer–Tafel or Volmer–Heyrovsky reactions. First, the surface-adsorbed hydrogen is rapidly formed by the combination of protons and electrons on the electrode (Volmer reaction, H+ + e− → H*).50 Subsequently, two optional subsequent steps: where one proton in water binds with another adsorbed H atom on the surface to form a H2 molecule (Heyrovsky reaction, H+ + e− + H* → H2),51 or Tafel reaction, which refers to two adsorbed hydrogen binding to each other to form a H2 molecule (2H* → H2).52
The potential barrier and temperature determine the reaction rate, so we calculated the transition state of VP-MA2Z4 using the CI-NEB method, as shown in Fig. 8. For VP-MoGe2P4, the potential barriers to the Tafel reaction and the Heyrovsky reaction are 3.31 eV and 1.05 eV, and the final state energy of the Heyrovsky reaction is −3.82 eV relative to the initial state. This means that the hydrogen evolution reaction of VP-MoGe2P4 is more suitable for the Heyrovsky reaction with a low potential barrier and the reaction is exothermic. Similarly, VP-WGe2P4 tends to have a Tafel reaction with a potential barrier of 2.95 eV, and VP-VGe2P4 tends to have a Heyrovsky reaction with a potential barrier of 2.27 eV. In summary, the VP-MoGe2P4 structure has the best catalytic performance and the lowest reaction barrier and is likely to carry out the Volmer–Heyrovsky reactions in solution.
|
| Fig. 8 The reaction pathways for the HER. (a) Tafel reaction of VP-MoGe2P4, (b) Heyrovsky reaction of VP-MoGe2P4, (c) Tafel reaction of VP-WGe2P4, (d) Heyrovsky reaction of VP-WGe2P4, (e) Tafel reaction of VP-VGe2P4 and (f) Heyrovsky reaction of VP-VGe2P4. The initial state (IS), the transition state (TS) and the final state (FS) are linked by dotted lines in the figure; the red line indicates the corresponding barrier. | |
4. Conclusion
We theoretically designed a defective MA2Z4 monolayer family, and investigated their stability and unique role as electrocatalysts toward the HER systematically. We find that VP-MA2Z4 possesses a superior HER performance over pristine MA2Z4. Importantly, the optimal HER activity can be achieved with VP vacancies active site 1, which indicates that the catalytic properties of the defect MA2Z4 can be tuned easily and effectively. Our calculations reveal that ΔGH* decreases significantly with the outer Z atom (VZ), so the vacancy defect at the outer Z atom is the most suitable regulation method for MA2Z4 materials. The electronic structure analysis shows that when vacancies are introduced, several new defect states move closer to the Fermi level, leading to semiconductor properties and an improvement of the hydrogen adsorption strength. We also find the charge transfer from the Z atoms to H* atom by calculating the electron charge density differences and Bader charges analysis, which indicates the effective performance of a HER catalyst. We further demonstrate that the HER on defect MA2Z4 prefers the Volmer–Heyrovsky mechanism. Our study shows that the designed defect MA2Z4 monolayer family is highly activated toward the HER electrocatalyst, the optimal HER activity can be achieved, and abundant catalytic activity sites are provided; MoGe2P4 with a P vacancy and outside P as the H adsorption site is found to be the best. It is expected that the strategies developed in this study may be applied for designing 2D electrocatalysts for low-cost and high-performance HER applications.
Conflicts of interest
There are no conflicts to declare.
Acknowledgements
This work was supported by the National Key Research and Development Program of China (No. 2021YFB3601201), the Open-Foundation of Key Laboratory of Laser Device Technology, China North Industries Group Corporation Limited (No. KLLDT202103) and the Open Project Program of Shandong Semiconductor Materials and Optoelectronic Information Technology Innovation Center, Ludong University. We are grateful for the helpful discussion with Prof. Pengfei Guan and the computational support from the Beijing Computational Science Research Center (CSRC).
Notes and references
- M. H. Wu, W. J. Chou, J. S. Huang, D. B. Putungan and S. H. Lin, Phys. Chem. Chem. Phys., 2019, 21, 21561–21567 RSC.
- G. Q. Yu, B. Y. Huang, X. Chen, D. Wang, F. Zheng and X. B. Li, J. Phys. Chem. C, 2019, 123, 21878–21887 CrossRef CAS.
- G. Zhang, Y. S. Feng, W. T. Lu, D. He, C. Y. Wang, Y. K. Li, X. Y. Wang and F. F. Cao, ACS Catal., 2018, 8, 5431–5441 CrossRef CAS.
- C. Yang, Z. Zhao and Q. Liu, Appl. Surf. Sci., 2022, 577, 151916 CrossRef CAS.
- J. A. Turner, Science, 2004, 305, 972–974 CrossRef CAS PubMed.
- P. C. K. Vesborg and T. F. Jaramillo, RSC Adv., 2012, 2, 7933 RSC.
- W. Qian, Z. Chen, J. Zhang and L. Yin, J. Mater. Sci. Technol., 2022, 99, 215–222 CrossRef.
- X. Qin, L. Dai, H. Li, K. Qu and R. Li, Int. J. Hydrogen Energy, 2022, 47, 15775–15782 CrossRef CAS.
- H. Vrubel and X. Hu, Angew. Chem., Int. Ed., 2012, 51, 12703–12706 CrossRef CAS PubMed.
- R. Michalsky, Y. J. Zhang and A. A. Peterson, ACS Catal., 2014, 4, 1274–1278 CrossRef CAS.
- W. Zhou, J. Jia, J. Lu, L. Yang, D. Hou, G. Li and S. Chen, Nano Energy, 2016, 28, 29–43 CrossRef CAS.
- W. Han, L. Chen, B. Ma, J. Wang, W. Song, X. Fan, Y. Li, F. Zhang and W. Peng, J. Mater. Chem. A, 2019, 7, 4734–4743 RSC.
- A. Wang, L. Shen, M. Zhao, J. Wang, W. Zhou, W. Li, Y. Feng and H. Liu, J. Mater. Chem. C, 2019, 7, 8868–8873 RSC.
- Z. Zhuang, Y. Li, Z. Li, F. Lv, Z. Lang, K. Zhao, L. Zhou, L. Moskaleva, S. Guo and L. Mai, Angew. Chem., 2018, 130, 505–509 CrossRef.
- J. F. Callejas, C. G. Read, E. J. Popczun, J. M. McEnaney and R. E. Schaak, Chem. Mater., 2015, 27, 3769–3774 CrossRef CAS.
- Y. Pan, K. Sun, Y. Lin, X. Cao, Y. Cheng, S. Liu, L. Zeng, W.-C. Cheong, D. Zhao, K. Wu, Z. Liu, Y. Liu, D. Wang, Q. Peng, C. Chen and Y. Li, Nano Energy, 2019, 56, 411–419 CrossRef CAS.
- H. Pan, Sci. Rep., 2015, 4, 5348 CrossRef PubMed.
- X. Fan, S. Wang, Y. An and W. Lau, J. Phys. Chem. C, 2016, 120, 1623–1632 CrossRef CAS.
- S. Chen, X. Chen, G. Wang, L. Liu, Q. He, X. B. Li and N. Cui, Chem. Mater., 2018, 30, 5404–5411 CrossRef CAS.
- Y.-J. Song and Z.-Y. Yuan, Electrochim. Acta, 2017, 246, 536–543 CrossRef CAS.
- B. Ren, D. Li, Q. Jin, H. Cui and C. Wang, J. Mater. Chem. A, 2017, 5, 19072–19078 RSC.
- J. Xiong, W. Cai, W. Shi, X. Zhang, J. Li, Z. Yang, L. Feng and H. Cheng, J. Mater. Chem. A, 2017, 5, 24193–24198 RSC.
- L. Li, R. Huang, X. Cao and Y. Wen, J. Mater. Chem. A, 2020, 8, 19319–19327 RSC.
- H. W. Kim, M. B. Ross, N. Kornienko, L. Zhang, J. Guo, P. Yang and B. D. McCloskey, Nat. Catal., 2018, 1, 282–290 CrossRef.
- G. Hussain, M. Asghar, M. Waqas Iqbal, H. Ullah and C. Autieri, Appl. Surf. Sci., 2022, 590, 153131 CrossRef CAS.
- X. Guo and S. Guo, J. Semicond., 2021, 42, 122002 CrossRef CAS.
-
L. Wang, Y. Jiang, J. Liu, S. Zhang, J. Li, P. Liu, Y. Sun, H. Weng and X. Q. Chen, arXiv, May 4, 2022, preprint, arXiv: 2205.01994, DOI:10.48550/arXiv.2205.01994.
- Y. Li and Y. Liu, New J.Phys., 2022, 24, 083008 CrossRef.
- Y. L. Hong, Z. Liu, L. Wang, T. Zhou, W. Ma, C. Xu, S. Feng, L. Chen, M. L. Chen, D. M. Sun, X. Q. Chen, H. M. Cheng and W. Ren, Science, 2020, 369, 670–674 CrossRef CAS PubMed.
- K. S. Novoselov, Natl. Sci. Rev., 2020, 7, 1842–1844 CrossRef PubMed.
- M. A. Abdelati, A. A. Maarouf and M. M. Fadlallah, Phys. Chem. Chem. Phys., 2022, 24, 3035 RSC.
- G. Kresse and J. Furthmüller, Phys. Rev. B: Condens. Matter Mater. Phys., 1996, 54, 11169–11186 CrossRef CAS PubMed.
- G. Kresse and D. Joubert, Phys. Rev. B: Condens. Matter Mater. Phys., 1999, 59, 1758–1775 CrossRef CAS.
- J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865–3868 CrossRef CAS PubMed.
- H. J. Monkhorst and J. D. Pack, Phys. Rev. B: Solid State, 1976, 13, 5188–5192 CrossRef.
- W. Tang, E. Sanville and G. Henkelman, J. Phys.: Condens. Matter, 2009, 21, 084204 CrossRef CAS PubMed.
- G. Henkelman, B. P. Uberuaga and H. Jónsson, J. Chem. Phys., 2000, 113, 9901–9904 CrossRef CAS.
- K. Tibbetts, C. R. Miranda, Y. S. Meng and G. Ceder, Chem. Mater., 2007, 19, 5302–5308 CrossRef CAS.
- B. Peng, F. Cheng, Z. Tao and J. Chen, J. Chem. Phys., 2010, 133, 034701 CrossRef PubMed.
- S. P. Kaur and T. J. D. Kumar, Appl. Surf. Sci., 2021, 552, 149146 CrossRef CAS.
- G. Liu, J. Li, C. Dong, L. Wu, D. Liang, H. Cao and P. Lu, Int. J. Hydrogen Energy, 2021, 46, 18294–18304 CrossRef CAS.
- Y. Zang, Q. Wu, W. Du, Y. Dai, B. Huang and Y. Ma, Phys. Rev. Mater., 2021, 5, 045801 CrossRef CAS.
- A. Bafekry, M. Faraji, D. M. Hoat, M. M. Fadlallah, M. Shahrokhi, F. Shojaei, D. Gogova and M. Ghergherehchi, J. Phys. D: Appl. Phys., 2021, 54, 155303 CrossRef CAS.
- J. Chen and Q. Tang, Chem.–Eur. J., 2021, 27, 9925–9933 CrossRef CAS PubMed.
- C. Xiao, R. Sa, Z. Cui, S. Gao, W. Du, X. Sun, X. Zhang, Q. Li and Z. Ma, Appl. Surf. Sci., 2021, 563, 150388 CrossRef CAS.
- J. S. Yang, L. Zhao, S. Q. Li, H. Liu, L. Wang, M. Chen, J. Gao and J. Zhao, Nanoscale, 2021, 13, 5479–5488 RSC.
- S. D. Guo, Y. T. Zhu and W. Q. Mu, Europhys. Lett., 2020, 132, 57002 CrossRef CAS.
- S. D. Guo, W. Q. Mu, Y. T. Zhu and X. Q. Chen, Phys. Chem. Chem. Phys., 2020, 22, 28359–28364 RSC.
- L. Wang, Y. Shi, M. Liu, A. Zhang, Y. L. Hong, R. Li, Q. Gao, M. Chen, W. Ren, H.-M. Cheng, Y. Li and X. Q. Chen, Nat. Commun., 2021, 12, 2361 CrossRef CAS PubMed.
- T. Erdey-Grúz and M. Volmer, Z. Phys. Chem., Abt. A, 1930, 150, 203–213 Search PubMed.
- J. Heyrovský, Recl. Trav. Chim. Pays-Bas, 2010, 46, 582–585 CrossRef.
- J. Wei, M. Zhou, A. Long, Y. Xue, H. Liao, C. Wei and Z. J. Xu, Nano-Micro Lett., 2018, 10, 75 CrossRef CAS PubMed.
|
This journal is © The Royal Society of Chemistry 2023 |
Click here to see how this site uses Cookies. View our privacy policy here.