Joy
Ekka
a,
Shrish Nath
Upadhyay
b,
Frerich J.
Keil
c and
Srimanta
Pakhira
*abd
aDepartment of Physics, Indian Institute of Technology Indore (IIT Indore), Simrol, Khandwa Road, Indore 453552, MP, India. E-mail: spakhira@iiti.ac.in; spakhirafsu@gmail.com
bDepartment of Metallurgy Engineering and Materials Science (MEMS), Indian Institute of Technology Indore (IIT Indore), Khandwa Road, Simrol, Indore 453552, MP, India
cDepartment of Chemical Reaction Engineering, Hamburg University of Technology, 21073 Hamburg, Germany
dCentre for Advanced Electronics (CAE), Indian Institute of Technology Indore (IIT Indore), Khandwa Road, Simrol, Indore 453552, MP, India
First published on 5th November 2021
Two-dimensional (2D) monolayer pristine MoS2 transition metal dichalcogenide (TMD) is the most studied material because of its potential applications as nonprecious electrocatalyst for the hydrogen evolution reaction (HER). Previous studies have shown that the basal planes of 2D MoS2 are catalytically inert, and hence it cannot be used directly in desired applications such as electrochemical HER in industry. Here, we thoroughly studied a defect-engineered Mn-doped 2D monolayer MoS2 (Mn–MoS2) material, where Mn was doped in pristine MoS2 to activate its inert basal planes. Using the density functional theory (DFT) method, we performed rigorous inspection of the electronic structures and properties of the 2D monolayer Mn–MoS2 as a promising alternative to noble metal-free catalyst for effective HER. A periodic 2D slab of monolayer Mn–MoS2 was created to study the electronic properties (such as band gap, band structures and total density of states (DOS)) and the reaction pathways occurring on the surface of this material. The detailed HER mechanism was explored by creating an Mn1Mo9S21 non-periodic finite molecular cluster model system using the M06-L DFT method including solvation effects to determine the reaction barriers and kinetics. Our study revealed that the 2D Mn–MoS2 follows the most favorable Volmer–Heyrovsky reaction mechanism with a very low energy barrier during H2 evolution. It was found that the change in the free energy barrier (ΔG) during the H˙-migration (i.e., Volmer) and Heyrovsky reactions is about 10.34–10.79 kcal mol−1 (computed in the solvent phase), indicating that this material is an exceptional electrocatalyst for the HER. The Tafel slope (y) was lower in the case of the 2D monolayer Mn–MoS2 material due to the overlap of the s-orbital of hydrogen and d-orbitals of the Mn atoms in the HOMO and LUMO transition states (TS1 and TS2) of both the Volmer and Heyrovsky reaction steps, respectively. The better stabilization of the atomic orbitals in the HER rate-limiting step Heyrovsky TS2 is the key for reducing the reaction barrier, and thus the overall catalysis, indicating a better electrocatalytic performance for H2 evolution. This study focused on designing low-cost and efficient electrocatalysts for the HER using earth abundant transition metal dichalcogenides (TMDs) and decreasing the activation energy barriers by scrutinizing the kinetics of the reaction to achieve high reactivity.
However, the above-mentioned reactions are competitive processes and are dependent on the electronic structure of the electrode surface. Thus, boosting the efficiency of the sluggish HER process is challenging. To date, platinum (Pt) is the best-known electrocatalyst for the HER due to its zero overpotential in acidic electrolytes.6 Because of their optimum Gibbs free energy (G) for the adsorption of atomic hydrogen, binding energy, and desorption of hydrogen from their surface with low activation energies, Pt-based catalysts have been long known as effective HER electrocatalysts.7,8 However, the limited availability and high cost of platinum limit its use as an efficient catalyst in large-scale commercial and industrial applications. To produce hydrogen on a global scale, we need to reduce the cost of its production. Thus, reducing the dependency on noble metal-based catalysts or even replacing them completely with non-noble metal alternatives would be a driving step towards a hydrogen economy. Accordingly, the search for Pt-free catalysts for the HER is of paramount importance in modern materials science and technology.
In the past few decades, countless exhaustive studies have focused on the development of earth abundant and ecofriendly materials showing excellent electrocatalytic effects for the efficient production of H2 through the HER. Consequently, many of them such as Pt, Au, and Pd noble metal-based electrocatalysts, nano-porous materials, two-dimensional transition metal dichalcogenides (2D TMDs), 2D–3D material alloys and 2D TMDs doped with other atom species have a significant impact as efficient HER catalysts.9–18 A plethora of breakthroughs has been achieved in this regard for the rational design of HER electrocatalysts. Recently, earth abundant layered TMDs, for example MoS2, have attracted tremendous interest due to their inherent properties, which produce hydrogen at a very low overpotential with a high current density.15,19,20 It is widely recognized that 2D monolayer MoS2 material can exist in several possible structures such as 1T (octahedral), 1T′ (distorted octahedral) and 2H (hexagonal) structures.21 The 2H phase structure (i.e., 2D monolayer 2H-MoS2) is the most stable21 and is commonly used for electrocatalysis of the HER.22–24 The high activity of 2D monolayer MoS2 is attributed to its appropriate Gibbs free energy of adsorption for atomic hydrogen.25 Other MoS2-based materials such as MoS2 di-anionic surface with controlled molecular substitution of their S sites with –OH functional group have also proven to be efficient electrocatalysts.26 However, although these materials show promising aspects, they are still not sufficient in their present form for large-scale industrial and commercial applications due to their inert basal plane. Theoretical calculations indicated that P-doped MoS2 shows good catalytic activity for the HER by reducing the change in free energy (ΔG). This is due to the P-doping in the pristine MoS2, which activates its inert basal plane;27 however, this non-metal doping is quite difficult because of the instantaneous formation of MoP.10 Furthermore, to avoid the formation of MoP, very expensive apparatus such as plasma ion implantation is required, and in some cases, special precautions have to be taken during P-doping in the MoS2.28 Thus, other methods and materials are required for developing low-cost and efficient HER catalysts to produce the desirable H2 for industrial and commercial applications. Several techniques have been developed to generate high-performance TMD-based materials, such as defect engineering, metal-atom doping, nanostructure engineering, interface and strain engineering, and phase engineering.15,29 However, another problem frequently encountered in the 2D TMDs is the stacking of their 2D layers, which decreases the number of exposed sites, and the conductivity along two stacked layers is extremely low,30 thereby impeding charge transfer and decreasing the HER performance of 2D TMDs. A prominent factor controlling the rate of the HER is conductivity; however, pristine 2D monolayer MoS2 shows semiconducting properties, indicating its low electrical conductivity, and thus is inadequate for large-scale commercial applications. One of the promising ways to enhance the HER activity of pristine 2D monolayer 2H-MoS2 is to expose its active sites.31 It was found that the most of the active sites of pristine 2D monolayer 2H-MoS2 for the HER are located at its Mo and S edge sites.32 Furthermore, to modulate its electron transport for achieving proper conducting pathways and enhance its hydrogen evolution activity, the doping of external elements in the pristine 2D monolayer 2H-MoS2 nanostructure is a promising way via modern technology.14 Therefore, mechanistic insights are important for designing efficient electrocatalysts for H2 evolution.
The development of operative, stable, and economic HER catalysts to overcome the challenges associated with H2 production via the electrolysis of water is significant in reducing production costs and extending the hydrogen economy. For example, engineering the HER activity of MoS2via co-confining selenium and cobalt on its surface and inner plane, respectively, has shown promising results.33 We propose that Mn-doping in the pristine 2D monolayer MoS2 TMD material can activate its inert basal plane and the Mn-doped 2D MoS2 (in short Mn–MoS2) can be a promising material for efficient H2 evolution.34,35 For transition metal-based catalysts, their performance is correlated with their surface electronic structures and the electronic configuration of the d-orbital of the transition metal.11 In this regard, we computationally developed a two-dimensional (2D) single-layer Mn-doped MoS2 material (i.e., Mn–MoS2) and investigated its electrocatalytic performance for efficient HER. Firstly, we performed first principles-based quantum mechanical (QM) hybrid periodic density functional theory (DFT)36–41 calculations to determine its electronic properties such as electronic band structures, band gap and total density of states (DOS). Recently, QM DFT approaches and molecular simulations have been employed for modeling heterogeneous catalytic reactions, adsorption, and chemical reactions on the surface of 2D metals.42–45 We found that the 2D monolayer Mn–MoS2 shows a zero band gap due to Mn-doping in the pristine 2D MoS2. The density of states calculation indicated that there is a large number of electronic states available around the Fermi energy (EF) level with a high availability of electrons due to the doping of Mn atoms in the pristine 2D monolayer MoS2 material.
One of the key features in determining the smooth flow of a reaction is the change in free energy (ΔG) of the possible reaction intermediates. Thus, to screen an appropriate candidate among the options available, it is important to compute the value of ΔG during hydrogen adsorption, which is an important parameter for evaluating the catalytic activity during the HER process. Lately, the quantum computational method has provided feasible procedures for calculating the free energy changes based on density functional theory (DFT).46–49 By modeling the possible reaction intermediates during the hydrogen evolution process on the surface of the electrocatalyst, its thermodynamic properties can be obtained using DFT methods. Therefore, we prepared a finite molecular cluster model system of the Mn–MoS2 material and carefully studied each reaction intermediate appearing during the HER process by employing the DFT method in both gas and solvent phase calculations. Our study showed that the 2D monolayer Mn–MoS2 TMD shows excellent catalytic activity for H2 evolution.
Fig. 1 Non-periodic finite molecular cluster model (Mn1Mo9S21) as derived from the 2D Mn–MoS2 monolayer material (represented by the dotted triangle). |
The electronic band structures and total DOS calculations were performed on the equilibrium structures of the TMDs by employing the same DFT-D method. All the integrations of the first Brillouin zone were sampled on 20 × 20 × 1 Monkhorst–Pack74k-mesh grids for the pristine 2D MoS2 and 4 × 4 × 1 for 2D Mn–MoS2. The k-vector path taken for plotting the band structure was selected as Γ–M–K–Γ for both materials (i.e., pristine MoS2 and Mn–MoS2). The atomic orbitals of Mo, S, and Mn were used to compute and plot the total DOS for the α electrons, which is enough to describe the electronic properties of the 2D Mn–MoS2 material. The single-point calculation was performed at the equilibrium geometry to form the normalized wave function at zero Kelvin temperature with respect to vacuum. To create the graphics and analysis of the 2D layer structures studied here, the visualization software VESTA75 was used. We are aware that the lateral interactions of the adsorbed species may change the free energies for different surface coverages. The same applies for different temperatures.42,44,45
The two horizontal dashed lines indicate terminations along the (100) Mn-/Mo-edge and (010) S-edge. The two triangles represent the terminations for the Mn-/Mo-edge and S-edge clusters and the dangling bonds in the finite cluster were set by considering a triangle, as shown in Fig. 2. Each Mo atom in the (001) basal plane of the finite molecular cluster model has an oxidation state of 4+ (and the oxidation state of the Mn atom is 4+) and they are bonded with 6 S atoms (3 S at the upper plane and 3 S at the lower plane of Mo), which gives a contribution of 4/6 = 2/3 electrons towards each Mo–S bonding, resulting in a stabilized structure. The same can be understood with the oxidation state of S in the basal plane. The sulfur (S) atom has an oxidation state of −2 and bonds with 3 Mo atoms, which results in a contribution of 2/3 electrons towards each Mo–S bond. Similarly, the edges of the (000) periodic molecular cluster is stabilized with 2 local electron Mo–S bonds (as well as Mn–S bonds) having a single electron contribution towards each bond. Thus, at the edges each Mo atom contributes 2 × 1 electrons towards the local Mo–S bonds plus 4 × (2/3) electron contribution towards 4 Mo–S bonding in the basal plane, as shown in the Fig. 1. This 14/3 {i.e., (2 × 1) + [4 × (2/3)]} contribution of electrons towards the Mo–S bonding of the edge Mo atom is satisfied with the d2 configuration of one Mo atom and d1 configuration of two Mo atoms at the edges. This configuration leads to a molecular system with the periodicity of 3, which results in the achievement of a stabilized molecular cluster model having three edges without any unsatisfied valency.88
To model the solvation effects, we used the polarizable continuum model89 (PCM) with water as the solvent. The PCM method using GAUSSIAN1647 uses an external iteration method, where the energy in the solution is computed by making the solvent reaction field self-consistent with the solute electrostatic potential, given that the real reaction occurs in the solvent phase. For specifying the molecular cavity used in PCM, we used the UAHF set, i.e., the United Atom Topological Model.90 We modeled our reaction mechanism in water having a static dielectric constant of 78.36 (zero-frequency, 298.15 K, 1 atm).90,91 The geometric optimization and molecular energy in the solvent phase were calculated using the above-mentioned method.
(1) |
ΔG = ∑GProducts − ∑GReactants | (2) |
ΔH = ∑HProducts − ∑HReactants | (3) |
ΔE = ∑EProducts − ∑EReactants | (4) |
System | Lattice constants (a = b) | Interfacial angles (α, β and γ) | Space group symmetry | Average bond distance | |
---|---|---|---|---|---|
Mo–S | Mn–S | ||||
MoS2 | 3.180 Å | α = β = 90.0° and γ = 120.0° | Pm2 | 2.411 Å | — |
Mn–MoS2 (3 × 3 × 1 supercell) | 9.451 Å | α = β = 90.0° and γ = 119.9 ° | P1 | 2.409 Å | 2.303 Å |
Here, it is confirmed from the electronic structure and properties calculations that the 2D monolayer Mn–MoS2 TMD material shows conducting properties (i.e., conducing in nature), and thus we proceeded with our study in the direction of theoretical and computational development of the optimum electrocatalyst. Hydrogen production from the electrolysis of water is conceived as suitable for the growing need for alternative green energy sources. In this regard, we carefully audit the fundamentals of the HER and control parameters of the kinetics of the reaction. We outline molecular simulation approaches that will help us to tackle the challenges in the design of cheap and practical catalysts.
Generally, it has been found that the Volmer–Heyrovsky reaction mechanism is more likely to be predominant when transition metal-based catalysts are used because of their good adsorption free energy.94 Thus, the proposed catalyst was analyzed for its involvement in the mechanics and kinetics of the reaction (as mentioned in Fig. 3). To study the HER mechanism, we computationally developed a cluster model system for the 2D monolayer Mn–MoS2 and performed non-periodic M06-L DFT theory. The two processes were carefully studied and discussed further.
For the evolution of an H2 molecule, protons and electrons must be added to the 2D Mn–MoS2 molecular cluster Mn1Mo9S21. Here, it is useful to initially examine the most stable structures with each number of extra electrons and each number of extra protons to understand the free energy differences between the intermediate states, and ultimately find the lowest-barrier pathway. A detailed description of the HER process involved in the subject reaction is required to explain the electrochemistry. The first process is the dissociation of water in the Heyrovsky reaction step. To initiate the HER, one electron is absorbed on the surface of the 2D monolayer Mn–MoS2 TMD, as depicted in Fig. 4. This step takes place under the SHE conditions, which is the basis of the thermodynamic potentials for oxidation and reduction processes. The first reduction is achieved with a reduction potential, ΔG, about −129.66 mV, resulting in the formation of [Mn–MoS2]− from [Mn–MoS2] with the changes in enthalpy (ΔH) and electronic energy (ΔE) of about −2.93 and −2.97 kcal mol−1, respectively, as shown in Table 2. Thereafter, H+ from the solvent medium is adsorbed on the sulfur site, which is the most energetically conducive site at that time, forming an [Mn–MoS2]Hs solvated cluster (as the first adsorption of H at the Mn site has an energy cost of ΔG = 3.57 kcal mol−1, the lower barrier path is to follow [Mn–MoS2]− → [Mn–MoS2]Hs rather than the [Mn–MoS2]− → [Mn–MoS2]HMn path). In the following step, the [Mn–MoS2]Hs−1 complex is formed due to the addition of another electron from the solvent with a second reduction potential of −199.91 mV. Hence, we report this HER as a two-electron transfer reaction. In the next step, the hydride (H−) ion from the sulfur site migrates to the neighboring responsive Mn site. The migration of H˙ from the S to Mn site is the H˙-migration Volmer reaction step. This transition structure, i.e., the Volmer transition state (TS) or H˙-migration reaction TS (TS1), is corroborated by the detection of an imaginary vibrational frequency at the site of the transition of hydride ion from S to Mn. The formation of the TS was accompanied by a positive free energy change of 7.23 kcal mol−1 in the gas phase calculations. H+ from the medium again attacks, and from here either the Tafel or the Heyrovsky process can take place. The Heyrovsky part is further shown in Fig. 4. Computationally, we explicitly added a 4H2O–H+ cluster (i.e., specifically 3 water molecules and one hydronium ion (H3O+)) in the vicinity of the active site of the catalyst and observed according to the simulation that the H− from the metal site (here Mn) and the H+ from the water cluster (i.e., adjacent hydronium ion) create a bond, and then evolve as H2. This process is called the Heyrovsky process, which is often mentioned as the Heyrovsky transition state (TS2) in the reaction mechanism during HER. The Mn–S bond distance during TS2 was recorded to be 2.311 Å, which is higher than the strain free bond distance of 2.303 Å, as mentioned in Table 1. Finally, H+(H3O+) splits off to form a neutral Mn–MoS2 surface. Alternatively, an electron is absorbed to obtain the neutral [Mn–MoS2]Hs. All the optimized reaction intermediates, reactants and products with the TSs of our proposed reaction scheme are illustrated in Fig. 5.
Reaction intermediates | ΔE (kcal mol−1) | ΔH (kcal mol−1) | ΔG (kcal mol−1) | |
---|---|---|---|---|
Step 1 | [Mn–MoS2] → [Mn–MoS2]−1 | −2.93 | −2.97 | −2.99 |
Step 2 | [Mn–MoS2]−1 → [Mn–MoS2]Hs | −26.98 | −27.21 | −27.25 |
Step 3 | [Mn–MoS2]Hs → [Mn–MoS2]Hs−1 | −2.48 | −2.91 | −4.61 |
TS1 | [Mn–MoS2]Hs−1 → Volmer TS | 8.17 | 8.03 | 7.23 |
Step 4 | Volmer TS → [Mn–MoS2]HMn−1 | −36.36 | −36.41 | −36.49 |
Step 5 | [Mn–MoS2]HMn−1 → [Mn–MoS2]HsHMn | −34.67 | −35.00 | −35.19 |
Step 6 | [Mn–MoS2]HsHMn → [Mn–MoS2]HsHMn + 4H2O_H+ | 0.04 | 0.02 | 0.01 |
TS2 | [Mn–MoS2]HsHMn4H2O_H+ → Heyrovsky TS | 10.69 | 10.63 | 10.59 |
Step 7 | Heyrovsky TS → [Mn–MoS2]Hs+1 | −28.20 | −28.83 | −29.38 |
We keenly followed the different reaction pathway schemes, as shown in Fig. 4, but kept our focus on the two important saddle points, i.e., the Volmer reaction where an H˙ atom migrates from S to the transition metal site (here Mn site), i.e., H˙-migration reaction TS1, and the other the Heyrovsky reaction step, where H+ from an adjacent water cluster and the H− from the Mn site recombine to form H2. It was observed that the Heyrovsky reaction step is the rate-determining step of the HER for our system of interest (i.e., 2D monolayer Mn–MoS2). The changes in free energy (ΔG), enthalpy (ΔH) and electronic energy (ΔE) at each reaction step during the HER process in the gas phase calculations are listed in Table 2. In summary, we observed that the Volmer activation barrier is about 7.23 kcal mol−1 and the H2 formation reaction barrier, i.e., Heyrovsky reaction barrier, is about 10.59 kcal mol−1 computed in the gas phase (as illustrated in Fig. 6). The values of the activation barriers for the respective steps in the gas and solvent phases, i.e., the values of ΔG during the formation of the TSs are summarized and listed in Table 3. The changes in enthalpy (ΔH) and electronic energy (ΔE) during the formation of TS1, i.e., H˙-migration TS, are 8.03 kcal mol−1 and 8.17 kcal mol−1, and similarly, the values of ΔH and ΔE during H2 formation, i.e., TS2, are 10.63 kcal mol−1 and 10.69 kcal mol−1, respectively, as reported in Table 2.
Fig. 6 Free energy diagram, i.e., potential energy surfaces (PES) of the Volmer–Heyrovsky reaction pathway on the surface of the 2D monolayer Mn–MoS2 material computed in the gas phase. |
Activation barrier | ΔG (kcal mol−1) in the gas phase | ΔE (kcal mol−1) in the solvent phase | ΔH (kcal mol−1) in the solvent phase | ΔG (kcal mol−1) in the solvent phase |
---|---|---|---|---|
Volmer reaction barrier | 7.23 | 11.84 | 9.60 | 10.34 |
Heyrovsky reaction barrier | 10.59 | 12.56 | 10.36 | 10.79 |
PCM calculations have been developed to incorporate the solvation effects during the HER. According to our present DFT calculations considering the PCM system, we predicted that the Volmer step energy barrier, i.e., the reaction barrier (ΔG) of TS1, is about 10.34 kcal mol−1 in the solvent phase with a change in electronic energy (ΔE) of about 11.84 kcal mol−1, as reported in Table 3. Recently, Yu et al.9 reported that the TS1 H˙-migration reaction energy barriers during the Volmer step in the case of 2D monolayer pristine MoS2, WS2 and the hybrid W0.4Mo0.6S2 TMD alloy are about 17.7 kcal mol−1, 18.1 kcal mol−1 and 11.9 kcal mol−1, respectively, computed in the solvent phase by considering PCM. Moreover, the Heyrovsky TS barriers for the pristine MoS2, WS2 and W0.4Mo0.6S2 TMD alloy were calculated to be 23.8 kcal mol−1, 21.3 kcal mol−1 and 13.3 kcal mol−1, respectively. In the case of the 2D monolayer Mn–MoS2 material, the calculated value of ΔG during the Heyrovsky reaction step, TS2, was 10.79 kcal mol−1 with a change of electronic energy of 12.56 kcal mol−1 in the solvent phase calculation. The changes in enthalpy (ΔH) during the formation of the TSs TS1 and TS2 computed in the solvent phase are about 9.60 and 10.36 kcal mol−1, respectively, obtained by using the M06-L DFT method. The energy barriers for the different materials are mentioned in Table 4. We reported from our DFT calculations that the proposed catalyst shows much lower activation energies during the HER, and thus the 2D monolayer Mn–MoS2 can be characterized as a highly efficient HER catalyst. Specifically, the present DFT calculations revealed that both the H˙-migration (TS1) and Heyrovsky reaction (TS2) barriers during the HER process on the active surfaces of the 2D monolayer Mn–MoS2 TMD material are the lowest compared to the other TMDs and their hybrid alloys, indicating an excellent electrocatalyst for effective H2 evolution. The potential energy surfaces (PES) of the Volmer–Heyrovsky reaction pathway for HER on the active surface of the 2D monolayer Mn–MoS2 material computed in the gas phase is shown in Fig. 6 with the values of ΔG during the reactions.
A comparison of different electrocatalysts based on their activation barriers in the reaction computed in the solvent phase can be visualized by the graphical illustration shown in Fig. 7. The figures depict that the 2D monolayer Mn–MoS2 TMD material shows the lowest activation barriers (i.e., both the Volmer and Heyrovsky reaction barriers) among the operational electrocatalysts mentioned in Table 4.
Fig. 7 Graphical illustration of (a) Volmer and (b) Heyrovsky reaction barrier in the solvent phase for different materials. |
According to the transition state theory95 (TST) or the activated complex theory by including DFT calculations, we determined the turnover frequency (TOF) for H2 evolution per edge of Mn atom in the 2D monolayer Mn–MoS2 catalyst. According to the theoretical determination, we used the formula: rate = (kBT/h) × exp(−ΔG/RT);88 where kB is the Boltzmann constant, T (here 298.15 K) is temperature, h is Planck's constant, R is the universal gas constant and ΔG corresponds to the free energy barrier. The TOF obtained for the 2D monolayer Mn–MoS2 from the H2 formation reaction energy barrier in the Heyrovsky mechanism (in solvent phase) is about 7.74 × 104 s−1. A high TOF value is suitable for efficient H2 evolution during the reaction. For example, the excellent and well-functioning electrocatalyst developed by Yu et al. (hybrid W0.4M0.6S2 alloy material) showed a TOF value as high as 1.1 × 103 s−1. The TOF values of other practical catalysts such has 2D monolayer MoS2 and WS2 are mentioned in Table 5 for comparison. The 2D Mn–MoS2 material showed a very high TOF value, supporting the fact that this material will show excellent and efficient performances during the HER.
We can observe that the 2D monolayer Mn–MoS2 shows comparable results to the 2D monolayer hybrid W0.4Mo0.6S2 alloy material. Therefore, it can be concluded that the 2D Mn–MoS2 is a better and practical alternative for superb catalytic performance in the HER. Another electrochemical parameter, i.e., the Tafel slope (y) (which gives information about the kinetics, rate-determining steps of the electrochemical reaction and the energy required to achieve activity), is also one of the important factors to assess the performance of electrocatalysts. The experimentally observed Tafel slope (y) for the 1T phase of MoS2 and WS2 is about 40 mV dec−1 and 55 mV dec−1, respectively (synthesized via lithium intercalation at room temperature).96,97 The Tafel slope of another efficient and functional catalyst, MoS2 nanoparticles grown on graphene, is 41 mV dec−1, where electrochemical desorption was the rate-limiting step during hydrogen catalysis.22 The Tafel slope for the hybrid W0.4Mo0.6S2 alloy material synthesized via the wet chemical route was reported to be 38.7 mV dec−1.9 The Tafel slope can also be calculated theoretically by considering the number of electrons transferred during the HER. As stated earlier, the proposed reaction is a two-electron transfer mechanism, and thus our DFT-D computed Tafel slope was 29.58 mV dec−1 for the 2D Mn–MoS2, which is 9.12 mV dec−1 lower than that of the hybrid W0.4Mo0.6S2 alloy material, indicating that the 2D Mn–MoS2 is an excellent electrocatalyst for the HER.
Our present computations are in strong favor of low energy barriers for both the Volmer and Heyrovsky steps in the HER using the 2D monolayer Mn–MoS2, which is a favorable candidate as an HER electrocatalyst. To further support our development, we implemented natural bond orbital (NBO), highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) calculations in the DFT analysis. These calculations were performed to show an appropriate perspective of H2 formation at the active site from an electronic charge and molecular orbital overlapping point of view. Precise Lewis structures, i.e., structures that have maximum electronic charge in their Lewis orbitals, can be found by calculating the NBO. This study conveyed the interaction density or the overlap density from the wavefunctions. The solution to a multielectron atomic system requires an approximation called the linear combination of atomic orbitals (LCAO approximation). The qualitative picture of the molecular orbital is analyzed by expanding the molecular orbital into a complete basis set of all the atomic orbitals of the nuclei. Thus, the multi-electron wavefunction in a molecule at a specific configuration of the nuclei can be given by expanding the orbital approximation to molecules. The wavefunction obtained from the NBO calculations is a linear combination of the atomic orbitals of the Mn, S, Mo and H atoms for the Volmer TS and the Mn, S, Mo, H and O atoms for the Heyrovsky TS. The HOMO–LUMO was obtained from the optimized transition structures (both Volmer and Heyrovsky transition states TS1 and TS2), as shown in Fig. 8. The red color represents the in-phase bonding of the orbitals and the blue color shows the out-of-phase bonding. The boundary value outlining the isosurface shown in Fig. 8 was set at 0.009. The interval of the values at which the isosurface is colored (from blue to red) was set from −0.1 to 0.1.
Insight can be gained into the role of the electronic structure in the HER mechanism from the HOMO calculations of the transition states and H2 formation in a steady state due to the better overlap of the d-orbital of the Mn atom and the s-orbital of H2 compared to the pristine 2D monolayer MoS2 or WS2. Therefore, one conclusion can be drawn here is that in the rate-limiting step of the HER, i.e., the Heyrovsky step, the stabilization of the atomic orbitals is also one of the key features for reducing this reaction barrier. The electron cloud around the H atoms in the Heyrovsky TS is highlighted by a red dotted circle (Fig. 8(c)). This step is supported by the overlap of the atomic orbitals of the H and Mn atoms together with H3O+ ion when H2 is evolved. This is also one of the reasons why the 2D monolayer Mn–MoS2 shows excellent activity for the HER. The energy difference between the HOMO and LUMO, also known as the HOMO–LUMO gap, is used to predict the stability of transition metal-based complexes98,99 given that it is the lowest energy electronic excitation that is possible in a molecule.
In this reaction, the adsorbed hydrogen at the sulfur site and the adsorbed hydrogen at the transition metal site combine to form H2 (2H* → H2, where H* represents hydrogen adsorbed on the active site of the catalyst). The Tafel reaction barrier (ΔG) in the gas phase was recorded to be 90.13 kcal mol−1 and 93.72 kcal mol−1 in the solvent phase, which are extremely higher than the Volmer–Heyrovsky reaction barrier. The changes in enthalpy (ΔH) and electronic energy (ΔE) during the formation of H2 in the Tafel reaction step are 92.27 kcal mol−1 and 93.66 kcal mol−1, respectively, computed by the DFT method, as reported in Table 6. The Tafel reaction barrier (ΔG) of TS3 is higher than the Heyrovsky reaction barrier TS2, indicating that the Volmer–Tafel reaction step is thermodynamically less favorable. HOMO–LUMO calculations were also performed to better visualize the Tafel reaction mechanism (Fig. 10(b) and (c), respectively). The electron cloud represents both positive and negative parts of the wavefunction in red and blue color, respectively. The electron cloud around hydrogen is highlighted by a red dotted circle. Fig. 10b presents the HOMO of the Tafel TS and the orbitals around H2 formed during the formation of the Tafel TS are highlighted by blue. This means that the orbital mixing is out of phase. The red cloud around H2 in the LUMO of the Tafel TS suggests the in-phase interaction of the electronic wavefunctions. The phase or orbital is a direct consequence of the wave-like property of electrons, where generally the in-phase mixing suggests a lower energy state and the out-of-phase mixing indicates anti-bonding orbitals or higher energy state. The corresponding TOF in both the gas and solvent phases was calculated to be 6.19 × 10−45 s−1 and 1.25 × 10−58 s−1, respectively. This TOF value is very low, and hence the process is less likely to occur.
Activation barrier | ΔG (kcal mol−1) in the gas phase | ΔE (kcal mol−1) in the solvent phase | ΔH (kcal mol−1) in the solvent phase | ΔG (kcal mol−1) in the solvent phase |
---|---|---|---|---|
Volmer reaction barrier | 7.23 | 11.84 | 9.60 | 10.34 |
Heyrovsky reaction barrier | 10.59 | 12.56 | 10.36 | 10.79 |
Tafel reaction barrier | 90.13 | 93.66 | 92.27 | 93.72 |
It is clear from the data (Table 6) that the Tafel barrier is much higher than the calculated Heyrovsky barrier, and hence the Volmer–Heyrovsky reaction will be more dominant than the Volmer–Tafel reaction when using 2D monolayer Mn–MoS2 material-based catalysts. Therefore, heteroatom doping in the pristine 2D monolayer MoS2 will lead to a significant change in the electronic properties of the material. As shown in our present computed results, the 2D monolayer Mn–MoS2 TMD material shows an excellent electrocatalytic performance. The results of the descriptor-based method aided by DFT computations were thoroughly discussed above. This indicates that 2D Mn–MoS2 driven catalysis is a viable and efficient hydrogen production method.
This journal is © the Owner Societies 2022 |