Yunji Han‡
ab,
Mihyeon Jo‡ab,
Hyung-Kyu Lim*c and
Sangheon Lee*ab
aDepartment of Chemical Engineering and Materials Science, Ewha Womans University, 52, Ewhayeodae-gil, Seodaemun-gu, Seoul 03760, Republic of Korea. E-mail: sang@ewha.ac.kr
bGraduate Program in System Health Science and Engineering, Ewha Womans University, 52, Ewhayeodae-gil, Seodaemun-gu, Seoul 03760, Republic of Korea
cDivision of Chemical Engineering and Bioengineering, Kangwon National University, Chuncheon, Gangwon-do 24341, Republic of Korea. E-mail: hklim@kangwon.ac.kr
First published on 13th August 2024
This theoretical investigation explores sustainable methods for ammonia synthesis by evaluating the effectiveness of dual-atom catalysts composed of transition metals anchored on boron α- and β12-sheet substrates for the electrochemical nitrogen reduction reaction. Utilizing density-functional theory, this study identifies enhancements in N2 adsorption capabilities, crucial for mitigating kinetic limitations in the electrochemical nitrogen reduction reaction. The research delves into the charge redistribution mechanisms between the catalyst and nitrogen, emphasizing the pivotal role of metal oxidation states and charge injection into nitrogen in influencing nitrogen binding and subsequent electrochemical reduction steps. Furthermore, this work defines optimal conditions for catalytic efficiency by identifying a specific range for the binding strength of reaction intermediates, ensuring the effectiveness of the Sabatier principle in terms of N2 binding. These insights significantly enhance the current understanding of catalyst design for the electrochemical nitrogen reduction reaction and provide a robust theoretical foundation for developing more efficient ammonia production systems.
The traditional method for ammonia synthesis, the Haber-Bosch process (N2 + 3H2 → 2NH3), is energy-intensive, requiring high temperatures (400–500 °C) and pressures (150–300 bar) to break the strong triple bond of a dinitrogen molecule (bond dissociation energy = 225 kcal mol−1).11,12 Despite its high energy consumption, accounting for about 2% of the global energy supply, and associated greenhouse gas emissions, this process remains the dominant industrial method for ammonia production. Accordingly, there is a growing emphasis on developing alternative ammonia synthesis routes that are both energy-efficient and devoid of fossil fuel dependence.
The electrochemical nitrogen reduction reaction (eNRR) (N2 + 6H+ + 6e− → 2NH3), which converts abundant atmospheric N2 to NH3, offers multiple advantages over the traditional Haber-Bosch process.13–16 Specifically, the eNRR operates under less stringent conditions, such as room temperature and atmospheric pressure, and has the potential to be powered by renewable electrical energy sources like wind and solar. This capability significantly reduces the overall carbon footprint for ammonia production. Although the equilibrium potential under aqueous conditions is +0.092 VRHE,17 complete conversion of N2 to NH3 under eNRR conditions requires a considerable overpotential due to the significant thermodynamic and kinetic barriers at each elementary step involved in the reductive addition of hydrogen to N2.18,19 While numerous experimental studies have been conducted to explore various types of materials – including pure metals,20,21 metal alloys,22–24 metal oxides,25,26 metal nitrides, and metal phosphides27,28 – the overpotential and faradaic efficiency have not yet reached levels suitable for commercial application. In this regard, it is necessary to overcome challenges like high overpotentials and low faradaic efficiencies by fine-tuning catalyst structures at the atomic level, based on a thorough understanding of the reactivity at each elementary step.
In recent years, single-atom catalysts (SACs) have received significant attention for their high catalytic efficiency, selectivity, and stability in various catalytic reactions.29–32 These catalysts are noteworthy because they allow for the more efficient use of precious metals and offer tunable active site configurations. Additionally, SACs provide a foundational framework for catalyst design, both experimentally and theoretically. Recently, the focus has also shifted toward dual-atom catalysts (DACs).33–41 These catalysts feature two closely adjacent metal atoms embedded in suitable substrates, such as N-doped carbons, graphitic carbon nitride (g-C3N4), or boron nitrides. DACs are thought to offer a synergistic effect, potentially improving both catalytic activity and durability compared to SACs. Notably, there has been a growing trend in computational active-site screening studies based on density-functional theory (DFT), especially in the field of the eNRR for SACs and DACs.36,42,43 Computational studies have demonstrated significant eNRR activity with various catalysts, including Fe–Rh anchored on nitrogen-doped graphene, showing an onset potential of −0.22 VCHE. Other notable examples are single V atoms anchored on boron β12-sheets with an onset potential of −0.28 VCHE and single B atoms supported on holey g-C3N4, which exhibited an exceptionally low overpotential of −0.15 VCHE.
Recently, boron sheets have emerged as candidate support materials for SACs or DACs, exhibiting 2D network structures with sp2 hybridization similar to graphene.44–47 These sheets are not only energetically stable but are also expected to have a range of chemical applications. Among the various boron sheet structures proposed, those with hexagonal holes inducing triangular lattice arrangements, such as α-sheets, and β12-sheets, have been theoretically confirmed as the most stable. In particular, the work by Er et al. shows that boron sheets facilitate easier metal atom adsorption than graphene due to their electron-deficient nature.48 Subsequently, the DFT study by Liu et al. suggests enhanced catalytic activity in the eNRR for Ru-doped boron α-sheet and β12-sheet SACs with low limiting potentials of ΔGmax = 0.42 and 0.44 eV for Ru on α- and β12-sheets, respectively.49 Despite the promising catalytic activities in terms of limiting potential, these SACs suffer from the problem that the non-electrochemical N2 adsorption step is thermodynamically unfavorable, so the overall eNRR rate is likely to be limited even if sufficiently high potentials are provided. These calculation results for Ru-doped boron sheet SACs indicate that for boron sheet catalysts to be competitive in the field of the eNRR, appropriate strategies must be jointly pursued to overcome the thermodynamically less favorable N2 adsorption step while minimizing the magnitude of the limiting potential. While various theoretical studies on DAC-based electrocatalyst screening have been conducted,33,50,51 research on the NRR using boron-sheet-based catalysts is still lacking. By addressing this unexplored area, our study aims to fill this gap in the literature and provides important insights into catalyst design for the NRR.
In this study, we perform additional DFT-based active-site screening for boron sheet eNRR catalysts by investigating the eNRR activity, when various transition metals (TMs) are doubly anchored onto boron α-sheets and β12-sheets in the form of a DAC. As a result, we gain insight into the design principles of metal-doped boron sheet catalysts applicable for overcoming the key kinetic limitations associated with the eNRR. Subsequent in-depth electronic structure analysis provides theoretical evidence on how electronic interactions between the doped metal and boron sheet regulate the eNRR activity in a positive direction. The insights from this theoretical work are expected to serve as a robust foundation for designing experimental approaches aimed at refining eNRR catalyst systems.
The adsorption energy (ΔEads,M) of a species M (M = N2 and its reaction intermediates) is calculated as the energy difference between the adsorption system (E*M), the gas-phase adsorbate molecule (EM), and the free-standing surface (E*), using the following equation:
ΔEads,M = E*M − E* − EM | (1) |
Note that the greater the ΔEads,M is along the negative direction, the stronger the adsorption strength of a species M is, and vice versa. The Gibbs free energy change (ΔG) for each elementary reaction step involving the proton coupled electron transfer (PCET) process is calculated using the computational hydrogen electrode (CHE) model.55 In the CHE model, the Gibbs free energy of (H+ + e−) is defined as the free energy of 1/2H2 in the gas-phase. Thus, the ΔG change is computed using the following equation:
ΔG = ΔEelec + ΔEvib − TΔSvib + ΔGsolv | (2) |
We perform crystal orbital Hamilton population (COHP) analysis using the Local-Orbital Basis Suite Towards Electronic-Structure Reconstruction (LOBSTER) package.57,58
The first step in the eNRR process is the adsorption of N2, which is one of the fundamental factors for determining the performance of a catalyst. Due to the inert nature of N2 molecules, they often fail to adsorb onto conventional catalyst surfaces, resulting in a kinetically unfavorable aspect for initiating the overall reaction.59 To determine the most stable N2 adsorption geometries on our DAC models, we optimize various possible N2 adsorption configurations with different orientations and symmetries, as illustrated in Fig. 1b. For the α-sheets, the distance between the active TM sites is relatively large, resulting in two adsorption modes: “end-on” where the N2 molecule adsorbs perpendicularly to a single active site, and “side-on” where the adsorbed N2 molecule is aligned parallel to the surface between two active sites. For the β12-sheet, the proximity of the active sites allows for two additional configurations, resulting in a total of four modes including “symmetric end-on” and “asymmetric side-on” modes in addition to “end-on” and “side-on” modes. It is important to note that the kinetic diameter of N2 (3.64 Å) is larger than the inter-vacancy distance in β12-sheets (2.98 Å). For chemisorption on catalyst surfaces, the actual bond length of N2 (1.01 Å) is more relevant to the reaction than the kinetic diameter of N2. The N2 bond length is significantly smaller than the inter-vacancy distance in the β12-sheet. Our DFT calculations show that upon metal atom adsorption, the distance between metal atoms in β12-sheets is long enough to accommodate N2 molecules in various configurations. Our detailed analyses of the possible N2 binding configurations, considering the actual molecular dimensions and the structural characteristics of our catalyst system, confirm the feasibility of all binding modes presented in Fig. 1b.
Fig. 1c and d show the most stable N2 adsorption energies and structures for each dual-atom TM embedded in boron α- and β12-sheets, respectively. For the α-sheet, most TMs, except Ag, Co, Cu, Fe, and Pd, form stable adsorption structures in the side-on configuration with negative adsorption free energies. In particular, Cr, Mo, Nb, Tc, Ti, V, and Zr exhibit significantly strong N2 fixation effects with adsorption free energies ranging from −0.55 eV up to −1.75 eV. These calculation results indicate that despite the relatively long distance between the TM sites, the incorporation of dual-atom TMs within a boron α-sheet enables stable N2 chemisorption on the TMs in the side-on configuration. This can be attributed to the structural flexibility of the boron sheet, resulting in optimized N2 adsorption structures with reduced metal–metal distances, initially from 5.13 Å to 4.40–4.75 Å (Table S4 and Fig. S1†). For the β12-sheet, more complicated N2 adsorption characteristics, due to the ensemble effects between the two closely spaced active sites, result in a variety of N2 adsorption structures with often marginal energy differences between different configurations. Nevertheless, similar to the α-sheet, the majority of stable N2 adsorption structures on the β12-sheet are of the side-on configurations, and strong adsorption cases are found exclusively in the asymmetric side-on configuration. Note that the metal–metal distances of the β12-sheet are marginally affected by N2 adsorption compared to those of the α-sheet (Table S4†). This indicates that the arrangement of metal active sites in the β12-sheet is structurally highly conducive to N2 adsorption.
To further elucidate the electronic structural origins of this trend, we conduct additional analyses including partial density of states (pDOS) and crystal orbital Hamilton population (COHP) analyses (Fig. S2–S5†). The pDOS analysis reveals that there is a correlation between the N2 adsorption strength and the formation of bonding orbitals between metal d orbitals and N 2p orbitals near the Fermi level. As shown in Fig. 3a, we find that the metal–N2 binding strength tends to increase as the amount of charge transfer to N2 increases. Moreover, COHP analysis demonstrates that there is a strong correlation between the COHP-calculated metal–N2 binding strength and the N2 binding free energy, as shown in Fig. 3b. These results provide a more comprehensive understanding of the N2 activation mechanism, showing that the NRR thermodynamics of the TM active sites exhibit systematic variations from both classical charge transfer and orbital interaction perspectives. This interpretation of bonding characteristics from dual perspectives offers valuable insights into the rational design of more efficient NRR catalysts, highlighting the importance of electronic structure engineering for optimizing catalyst performance.
Fig. 3 (a) Calculated N2 adsorption free energies as a function of (a) the average partial charge of a N2 molecule adsorbed on DACs and (b) COHP-calculated metal–N2 binding strength on α-sheets (denoted as α) and β12-sheets (denoted as β) for the most stable N2 adsorption structures (see Fig. 1). Refer to raw data in Table S7.† |
These types of positively charged active sites are also known to be beneficial in suppressing competitive hydrogen evolution reactions due to the enhanced electrostatic repulsion with protons. One may wonder whether the positively charged active sites could be poisoned by H2O or OH adsorption. Our calculations of water molecule adsorption on these oxidized active sites revealed that, with the exception of group 4 and 5 metals such as Nb, Ti, V, and Zr (which have the highest oxidation states), water adsorption is either difficult or weak (Fig. S6 and Table S9†). This suggests that for most of the active sites, side reactions related to water molecules are unlikely to have a significant impact on the NRR process.
Further analysis using N2 adsorbed models shows that the adsorbed N2 tends to be negatively charged, with adsorption strength being proportional to the negative partial charge on N2 (Fig. 2c, d and Table S6†). This indicates that electrons are transferred to the N2 molecule, with the transfer reaching levels up to approximately one electron per molecule. This implies that boron-based DACs can overcome the challenges of electron injection, which are typically hindered by the strongly negative electron affinity (−1.9 eV) of the N2 molecule.63 It is worth noting that electrons taken up by N2 mostly transferred from the TM active site (Table S8 and Fig. S7†). This suggests that TM active sites on the boron surface have the capability to transfer additional electrons to the adsorbate. Therefore, a fundamental correlation is confirmed that the strength of N2 binding correlates with the oxidation tendencies of the TM species, which is supported by the positive correlation between the ionization energy of TMs and N2 binding energy (Fig. S8†). Considering the maximum N2 adsorption free energy of around −0.5 eV in a previous boron-based SAC study,49 the dual-atom ensemble formation of DACs enables tight side-on interactions with N2, thereby achieving more stable bonding with enhanced electron transfer.
Fig. 4 Correlation between N2 and N2H binding free energies for (a) α-sheets (denoted as α) and (b) β12-sheets (denoted as β). (c) Selected eNRR pathways depending on the N2 binding structures. |
The conventional eNRR pathways can be categorized into distal and enzymatic pathways, depending on the starting N2 binding structure (Fig. S11†).59 In the distal pathway, the first three PCETs for the formation of the first NH3 molecule occur on the distal nitrogen atom in end-on modes. Subsequently, the following three PCETs take place on the remaining single nitrogen atom to form another NH3 molecule. In the enzymatic pathway, due to the symmetric nature of the N2 binding in a side-on configuration, six consecutive PCETs take place alternately on both nitrogen atoms, resulting in the formation of two NH3 molecules in the final two steps. In the α-sheets, the large distance between active sites clearly differentiates between end-on and side-on N2 binding modes, leading to distinct distal (Ag, Co, Cu, Fe, and Ni) and enzymatic reaction pathways (rest of the TMs). In contrast, the β12-sheets feature shorter distances between active sites, resulting in minor variations in N2 binding energy depending on the structure. Consequently, the binding mode of reaction intermediates may change during consecutive PCET steps due to the ensemble effect of dual active sites. This suggests that the most feasible reaction pathway is not solely dependent on the initial N2 binding mode. Investigations into the possible binding modes of N2H and N2H2 identify the most stable configurations for each TM, leading to the conclusion that the eNRR pathway for the β12-sheet follows the conventional enzymatic and assisted distal pathways (Fig. 4c). As a result, every TM with end-on (Ag, Co, Cu, Fe, Ni, and Pd) and symmetric end-on (Ru) N2 binding stabilizes into a symmetric side-on N2H binding mode, thus adhering to the enzymatic pathway. Additionally, among the asymmetric side-on TMs, Cr, Mo, and Nb shift to the enzymatic pathway after the 1st and 2nd PCET steps, respectively, while the remaining TMs follow the assisted distal pathway.
Achieving optimal N2 binding on the catalyst surfaces significantly impacts the kinetics of the eNRR, particularly by lowering the energy barrier associated with the first PCET step. This adjustment is crucial because the initial PCET step has traditionally been a significant kinetic bottleneck in the eNRR process due to the inherently low proton affinity and negative electron affinity of the N2 molecule. By fine-tuning active sites of the catalyst to enhance N2 adsorption, the electron density around the N2 molecule can be effectively increased, facilitating easier proton attachment and reducing the energy required for this critical step. Consequently, this optimized N2 binding not only decreases the activation energy barrier but also enhances the overall reaction kinetics, making the eNRR process more efficient and potentially more competitive with traditional ammonia synthesis methods. Furthermore, to achieve ideal catalyst performance with equally distributed reaction energies across all PCET steps, future research should also focus on breaking the linear scaling relationships observed in current catalyst systems. This insight underscores the importance of designing catalysts with precise control over N2 adsorption characteristics to overcome the kinetic challenges and improve the efficiency of ammonia production via electrochemical methods.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ta04256e |
‡ The two authors contributed equally. |
This journal is © The Royal Society of Chemistry 2024 |