Download PDF
Research Article  |  Open Access  |  20 Dec 2023

Computational design of spatially confined triatomic catalysts for nitrogen reduction reaction

Views: 452 |  Downloads: 169 |  Cited:   0
J Mater Inf 2023;3:26.
10.20517/jmi.2023.35 |  © The Author(s) 2023.
Author Information
Article Notes
Cite This Article

Abstract

The electrocatalytic process of nitrogen reduction reactions (NRR) offers a promising approach towards achieving sustainable ammonia production, acting as an environmentally friendly replacement for the conventional Haber-Bosch method. Density functional theory calculations have been utilized to design and investigate a set of catalysts known as triple-atom catalysts (TACs) for electrochemical NRR, which are supported on graphite-C3N3 nanosheets. Herein, we have systematically evaluated these TACs using stringent screening to assess their catalytic performance. Among the candidates, supported Pt3, Re3, and Ru3 trimers emerged as highly active with decent selectivity, involving a limiting potential range of -0.35~-0.11 V. According to analysis of electronic properties, we determined that high NRR activity stems from the d-π* electron-accepting and -donating mechanism. Significantly, the correlation between chemical activity of TACs and electronic structure was established as a pivotal physical parameter, which has led to the conclusion that we can precisely control the catalytic behavior of transition metal trimer clusters by selecting appropriate metal elements and designing moderate cluster-substrates interactions. In summary, these theoretical studies not only enhance our understanding of how catalytic properties are governed by metal-support interactions, regulating stability, activity, and selectivity, but also offer a useful method for screening and designing novel TACs for NRR.

Keywords

Density theory calculation, triple-atom catalysts, nitrogen reduction reaction, metal-support interactions

INTRODUCTION

Ammonia (NH3) plays a pivotal role as a primary precursor for the production of chemical fertilizers, nitric acid, biofuel energy, plastic, synthetic fiber, etc.[1-4]. Industrially, NH3 is primarily synthesized through the Haber-Bosch process, involving the reaction of N2 and H2 at high temperatures (T > 700 K) and pressure conditions (P > 200 atm)[5-8]. However, this method consumes a substantial amount of global energy and contributes to significant greenhouse gas emissions[9,10]. Therefore, it is an urgent need to develop sustainable and clean methods for yielding NH3 products. Electrochemical nitrogen reduction reactions (NRR) utilizing renewable energy to convert N2 to NH3 have recently emerged as a promising alternative[11-18]. For instance, Geng et al. reported the excellent NRR activity of Ru on N-doped carbon, achieving a 29.6% Faradaic efficiency (FE) and an NH3 production rate of 120.9 μg·mg-1·h-1[19]. Additionally, the achieved NH3 yield rate reached 3.665 mg·h-1·mgRu-1 at a potential of -0.21 V, where the researchers successfully developed a novel approach using N-doped porous carbon (PC) to encapsulate single Ru sites for efficient NRR while maintaining the FE below 9%[20]. A 5,7-membered carbon ring-involved PC was developed for the electrocatalytic NRR of Ru-embedded PCs by Han et al.[21]. These materials, with an impressive NH3 yield rate as high as 67.8 ± 4.9 μg·h-1·mgcat-1 and a high FE of 19.5% ± 0.6%, exhibit remarkably favorable catalytic NRR properties as electrocatalysts, surpassing the majority of documented single-atom NRR catalysts. Moreover, theoretical research and calculation models play a key role in predicting highly active and selective catalytic materials[22], providing important reference values for catalyst preparation. According to density functional theory (DFT) calculations, Azofra et al. reported that V3C2 possesses the best NRR activity with a 0.64 eV activation barrier among d2-d4 M3C2 Mxenes[23]. Some theoretical work has studied the NRR catalyzed by TM@N4-G, in which the central transition metal (TM) atom is coordinated by four pyridinic nitrogen atoms. The results show that the limiting potentials of Ti@N4 (0.69 eV) and V@N4 (0.87 eV)[24] are shown to exhibit lower free energy for NRR than that of the Ru(0001) stepped surface (0.98 eV)[25]. Despite significant progress in this field, many obstacles remain, including high overpotentials (> 0.6 V) and low FEs (9%~29.6%) for NRR. Hence, the development of highly efficient and selective NRR electrocatalysts to facilitate mild condition synthesis of ammonia is crucial.

Over the last few years, the design and synthesis of catalysts have been revolutionized by the advent of atom dispersions[26,27]. Single-atom catalysts (SACs) have attracted attention for their high-specific activity and maximum metal utilization efficiency[28-32]. For example, He et al. found that a graphdiyne monolayer supported with 11 TM atoms (TM@GDY) exhibits exceptional stability as an electrocatalyst for hydrogen evolution reactions (HER) and oxygen evolution reactions (OER), involving an overpotential range of 0.01~0.46 V[33]. Moreover, Liu et al. conducted research on the electrocatalytic generation of NH3 from N2 at room temperature and atmospheric pressure, using nitrogen-doped PC embedded in cobalt, achieving a high ammonia generation rate of 0.86 μmol·cm-2·h-1[34]. Alternatively, other single metal atoms anchored on N-modified carbon-based materials, such as graphitic carbon nitride (g-C3N4) and defective graphene, are promising electrocatalysts for NRR with 0.34 V potential[35]. Despite the potential benefits of SACs, a significant challenge persists in balancing the reaction rate and FE for NH3 synthesis, mainly due to the involvement of multiple reactive species in the NRR. This challenge remains a major obstacle in the practical application of SACs for the controlled and efficient electrocatalytic generation of NH3.

Atomic clusters[36-38], on the other hand, possess unique geometric and electronic properties owing to their strong quantum size effects, rendering them highly promising for catalytic applications. The catalytic activities and product selectivity of supported metal nanoclusters (NCs) were governed with atomic precision by tuning the size and composition of metal NCs. Experimentally, the use of well-dispersed Pt2 dimers on graphene resulted in a unique catalytic system that significantly increased the aqueous hydrogenation rate of ammonia-borane in comparison to isolated individual Pt-ions. Specifically, the specific rate of the reaction was found to be approximately 17 times higher when using the Pt2 dimers[39]. Furthermore, a Mo3 trimer on graphdiyne nanosheets was observed to exhibit the highest activity, selectivity, and stability towards NRR based on the designed screening criteria, involving a calculated initial voltage of -0.32 V[40]. More importantly, the presence of several active sites in a catalyst is essential to broaden the range of adsorbate binding capabilities and to facilitate the catalysis of a broader spectrum of complex reactions to yield diverse products. Tian et al. fabricated diatomic Fe2 NCs anchored on mesoporous carbon nitride, which demonstrated remarkable catalytic activity in the conversion of substilbene to oxide, achieving a high selection of 93% while maintaining a transformation rate of 91%[41]. Ru3 stabilized on nitrogen-doped carbon nanosheets was reported to efficiently catalyze the selective oxidation of alcohols[42]. Moreover, theoretical studies have suggested that supported metal NCs exhibit remarkable abilities in catalyzing reactions involving the activation of inert reactant molecules or requiring multiple reaction centers[43-47]. For example, a tetramer immobilized in nitrogen-doped graphene monolayers shows optimal surface activities for N2 activation and subsequent reaction to yielding NH3 products, involving the calculational on-set potential of -0.45 V, significantly lower than that of supported single atoms and dimers[47]. Li et al. discovered that Rh3-C2N exhibited superior catalytic performance in the enzymatic pathway with a limiting voltage of only -0.45 V compared to Rh1-C2N and Rh2-C2N[48]. Furthermore, Zheng et al. have designed a set of electrocatalysts known as triple-atom catalysts (TACs). The theoretical insights suggest that the Co3-N4 system has a superior catalytic activity, attaining a boundary potential of -0.41 V via the enzymatic mechanism for NRR[46].

Motivated by the substantial progress in both experimental and theoretical studies, we have systematically investigated the stability of 3d-5d TM trimers embedded on C3N3 nanosheets (marked as TM3@C3N3) using DFT calculations and evaluated their electrocatalytic performances for NRR. The theoretical results indicate that TM3@C3N3 (TM = Re, Ru, Pt) holds significant promise electrocatalysts for the NRR, demonstrating excellent performance and feasibility. Significantly, the Re3@C3N3 system demonstrates exceptional catalytic activity, surpassing other catalysts, with a limiting potential of -0.11 V using the consecutive mechanism. Furthermore, we established a correlation between the inherent electronic characteristics of TM3@C3N3 and its catalytic performance, identifying significant physical parameters. Additionally, we advocate the concept of designing TACs as a strategic approach to drive the development of advanced electrocatalysts within the framework of green hydrogen economics.

MATERIALS AND METHODS

The Vienna ab initio simulation package (VASP) was utilized to conduct spin-polarized DFT computations, employing the plane wave basis set[49,50]. We employed projector augmented wave (PAW) potentials along with an energy cutoff of 500 eV[51,52]. The Perdew-Burke-Ernzerhof (PBE) functional of generalized gradient approximation (GGA) was employed as the exchange-correlation functional[53]. It has been extensively tested and successfully applied to an array of solid-state and molecular systems[22-26,54]. The van der Waals (vdW) interaction was investigated using Grimme’s semiempirical DFT-D3 scheme with dispersion correction[55,56]. The energy and force convergence thresholds were established at 10-4 eV and 0.02 eV·Å-1, correspondingly. To avoid any interaction between two periodic units, a (2 × 2) supercell of C3N3 monolayer with a 16 Å vacuum space along the z-direction was employed. Moreover, the Brillouin zone was sampled at the k-point using a 2 × 2 × 1 grid[57]. Ab initio molecular dynamics (AIMD) simulations were performed at a temperature of 300 K in an NVT ensemble for a duration of 10 ps in order to estimate the thermodynamic stability of the TM3@C3N3 systems.

The binding energies (Eb) were calculated to judge the thermodynamic stabilities of the designed TM3@C3N3 systems, as follows:

$$ E_{b} = E_{\rm TM_{3}@C_{3}N_{3}} - E_{\rm C_{3}N_{3}} - E_{\rm TM_{3}} $$

where $$ E_{\rm TM_{3}@C_{3}N_{3}} $$, $$ E_{\rm C_{3}N_{3}} $$, and $$ E_{\rm TM_{3}} $$ denote the total energy of C3N3 substrate anchored with transition-metal trimeric clusters, the optimized pristine C3N3, and the isolated transition-metal trimeric clusters, respectively. The electronic adsorption energy (Eads) of reaction intermediates on TM3@C3N3 substrates can be computed by the following formula:

$$ E_{\rm ads} = E_{\rm tot} - E_{\rm cat} - E_{\rm adsorbate} $$

where Etot is the total energy of TM3@C3N3 substrates adsorbed by the intermediate, and Ecat and Eadsorbate are the energy of TM3@C3N3 and the adsorbed intermediate, respectively. According to the proposal by Nørskov et al.[58-60], the calculated hydrogen electrode (CHE) model can be employed to determine the Gibbs free energy change (∆G) for each individual step in the electrochemical hydrogenation process, and the ∆G value is calculated by employing the formula as follows:

$$ \Delta G = \Delta E + \Delta E_{\rm ZPE} - T\Delta S + \Delta G_{U} + \Delta G_{\rm pH} $$

where ∆E is the total reaction energy gained from the DFT calculations. ∆EZPE and ∆S are the changes in zero-point energy and entropy, respectively. The zero-point energy and entropy were determined by calculating the vibrational frequencies. The temperature (T) was set to 298.15 K in this research. ∆GU denotes the impact of the applied potential (U) and is equal to -neU, where n corresponds to the number of electrons transferred. The correction of pH, ∆GpH, represents the free energy and can be calculated according to the formula: ∆GpH = -KBT × pH × ln10. The pH value is assumed to be 0, and KB refers to the Boltzmann constant. The quantities of transferred electrons and the potential of the applied electrode are represented by e and U, respectively. The potential determination step (PDS), which has the maximum ∆G value, can be used to determine the limiting potential (Ulimiting) of the entire reduction process in an acid solution using the following formula:

$$ U_{\text {limiting }}=-\Delta G_{\max } / e $$

RESULTS AND DISCUSSION

Geometry and stability of TM3@C3N3

Since the NRR is a complex process including various reactive species, the use of effective screening descriptors is crucial. As illustrated in Figure 1, we propose a standard strategy for screening candidate catalysts for N2 reduction reaction. (1) Thermodynamic stability: TACs should have thermodynamics (∆Eb < 0 eV, where ∆Eb is the binding energy of TM3 atoms on C3N3). Additionally, their dynamic stability should be confirmed through AIMD simulations at 300 K, ensuring that the structure remains stable without deformation; (2) Surficial activity: N2 should undergo complete activation ($$ \Delta G_{\rm \ast N_{2}} $$ < 0.50 eV, indicating chemical adsorption of N2, where the asterisk * denotes the adsorption site)[61]; (3) Energy cost: the potential range of NRR reaction on Ru-based catalysts with high catalytic performance under mild conditions is -0.50~-0.56 V. Therefore, in this paper, catalytic properties of NRR: the limit potential, Ulimiting, is used to evaluate NRR activity. To ensure low energy costs, we use a standard of 0.55 eV. For the first and last hydrogenation steps, which are typically the most likely limiting steps in a reaction, it is desirable to have G values that are as low as possible ($$ \Delta G_{\ast \rm{N_{2}}\rightarrow \text{NNH}} $$ ≤ 0.55 eV and $$ \Delta G_{\rm{NH_{2}}\rightarrow \rm{NH_{3}}} $$ ≤ 0.55 eV)[61,62]. The calculation of G is performed using Equation (3)[61]; (4) NRR vs. HER selectivity: to ensure high NRR selectivity, the Gibbs free energy value of hydrogen adsorption should exceed that of nitrogen. These criteria collectively form a robust filtration strategy for identifying potential NRR catalysts.

Computational design of spatially confined triatomic catalysts for nitrogen reduction reaction

Figure 1. The schematic diagram for screening candidate TACs for NRR. HER: Hydrogen evolution reactions; NRR: nitrogen reduction reactions; TACs: triple-atom catalysts.

Next, we first consider the stability of the metal trimer on C3N3 before delving into the electroreduction process. Through extensive geometric optimization and configuration search, we obtained the structures of TM3@C3N3. As summarized in Figure 2A, these metal trimer clusters show the strong binding strength on the C3N3 monolayer, involving Eb of -9.57~-4.22 eV, which indicates decent thermodynamic stabilities. The Eb is calculated by Equation (1) with the detailed data shown in Table 1. Furthermore, according to the above calculation, it is worth noting that Pd3@C3N3 possesses the weakest binding strength among the 21 selected systems. In light of this, we conducted AIMD at 300 K for 10 ps to evaluate the stability of TM3@C3N3, with Pd3@C3N3 considered as a representative. Figure 2B and C illustrates the oscillations of the DFT total energies (E) relative to the initial conditions and temperature (T), along with the dPd-N and dPd-Pd, indicating dynamic fluctuations near the initial condition. According to the captured structures of Pd3@C3N3 under different times [Figure 2D], it is preserved well and maintains structural stability, in which the vertical buckling exhibits minimal fluctuation, measuring less than 0.10 Å. Therefore, combined with the above calculations, we can confidently assert that these TM3@C3N3 systems have robust structural stabilities.

Computational design of spatially confined triatomic catalysts for nitrogen reduction reaction

Figure 2. (A) Computed the binding energies (ΔEbind) of triple-atoms anchored on C3N3. The grey, blue, and light green are C, N, and metal atoms, respectively; (B) Variations of the temperature (T) and total energies (E); and (C) the distances of Pd-N (dPd-N) and Pd-Pd (dPd-Pd) vs. time for AIMD simulations of Pd3@C3N3; (D) Pd3@C3N3 structures are captured within 10 ps, with the Pd shown in green. AIMD: Ab initio molecular dynamics.

Table 1

DFT-calculated the average distances of TM-N (dTM-N) and TM-TM (dTM-TM), vertical buckling between TM3 and C3N3 monolayer (dTM-C), combining energy (ΔEbind), d orbital centroid (εd), magnetic moment (Mag), and the number of CT between TM3 and C3N3 monolayers

SystemdTM-N (Å)dTM-TM (Å)hTM-C (Å)ΔEbind (eV) εd (eV)Mag (μB)CT (e)
Ti2.012.521.27-9.140.723.231.60
V2.032.381.53-9.010.460.000.58
Cr1.941.930.89-8.63-0.560.000.24
Mn1.972.330.97-7.44-1.494.960.60
Fe1.922.271.47-6.98-1.936.160.33
Co1.882.221.23-7.35-1.533.000.47
Ni1.862.211.17-7.47-1.580.000.46
Y2.183.532.01-9.481.870.991.27
Zr2.093.001.37-9.460.771.921.97
Nb2.072.591.31-7.950.770.771.63
Mo2.092.191.23-7.51-0.920.610.38
Ru2.032.511.73-7.38-1.590.000.30
Rh2.072.551.50-6.46-1.750.710.35
Pd2.112.591.37-4.22-2.090.000.24
Hf2.892.101.41-9.571.011.241.85
Ta2.552.101.28-9.070.390.071.00
W2.452.081.17-6.59-0.350.110.88
Re2.351.991.06-9.13-1.010.960.65
Os2.432.061.20-8.18-1.410.000.47
Ir2.422.020.68-7.70-2.020.000.40
Pt2.512.211.77-5.25-2.530.020.26

Activation of N2 on TM3@C3N3

In the overall electrochemical NRR process, the primary and critical step involves the adsorption and activation of N2 molecules, a process of utmost importance as it is responsible for activating the N≡N triple bond, laying the foundation for the smooth protonation to follow. Figure 3A shows that the activation of N2 is facilitated via an electron transfer mechanism, where the partially filled d orbitals of TM atoms accept electrons from N2 molecules while simultaneously donating d electrons to the anti-bonding orbitals (π*) of *N2 in the reverse direction. Therefore, this interaction strengthens the TM–N bond while attenuating N≡N. Notably, the metal double or triple-atom centers are critical to maximizing the activation effect by donating a greater number of electrons to N2 in comparison to monatomic active sites[63-66]. Using dual-metal or triple-atom centers proves to be a more effective approach in promoting N2 activation. Employing this approach provides an advantage in surmounting the barrier encountered during the initial step in the hydrogenation of NRR. By incorporating multiple metal atoms in the catalytic centers, the activation of N2 becomes more efficient, enhancing the overall performance of the NRR process.

Computational design of spatially confined triatomic catalysts for nitrogen reduction reaction

Figure 3. (A) Simplified schematic illustration of N2 binding to single- and double-atom sites; (B) Computed molecular orbitals showing the electronic structure of free N2 and N2 absorbed on Pt3/g-C3N3. The red dashed line is the d-band center; (C) The adsorption free energies ($$ \Delta G_{\ast \rm{N_{2}}} $$) of *N2 on TM3@C3N3. The number of CT of *N2 with (D) side-on and (E) end-on configurations vs. its adsorption energies ($$ \Delta E_{\ast \rm{N_{2}}} $$); (F) The d-orbital center (εd) of TM3@C3N3vs.$$ \Delta E_{\ast \rm{N_{2}}} $$. The inset is the differential charge density of N2 on TM3@C3N3 and TM3@C3N3. CT: Charge transfers; DOS: density of states; TM: transition metal.

Revealing the source of NRR activity in electrocatalysts provides guidance for designing and developing highly active catalysts. To uncover the fundamental mechanism behind the activation of N2, we employed the PBE functional to calculate the partial density of states (PDOS), enabling us to gain deeper insights into the nature of the interactions and the extent of N2 activation facilitated by the metal trimers[67,68]. The enhanced ability of Pt3/g-C3N3 to adsorb and activate N2, as compared to the free N2, can be attributed primarily to the presence of vacant and filled d orbitals. As displayed in Figure 3B, the empty 5d orbitals of Pt3 receive electrons from N2, resulting in the formation of bonding states. Simultaneously, a robust d-2π* orbital coupling takes place, causing the 2π* orbital from N2 to be fractionally filled in the vicinity of the Fermi level. The partial filling is accomplished by the electron feedback from the occupied d orbitals of Pt3 to the 2π* orbital of N2. The observed outcome provides additional evidence to support the presence of strong d-π* orbital coupling in close proximity to the Fermi level. Additionally, it affirms the significant overlap between the d orbitals of trimeric clusters and the filled orbitals of absorbed N2. This finding strongly aligns with the earlier analysis and supports the proposed mechanism of N2 activation.

To further understand the interacting nature of N2 and TM3@C3N3 catalysts, we conducted calculations to determine the $$ \Delta G_{\rm{N_{2}}\ast} $$ value. The corresponding adsorption free energy for N2 adsorption on TM3@C3N3 is calculated by Equation (3), with the detailed data presented in Figure 3C and Supplementary Table 1, respectively. It is not difficult to see that for most TM3@C3N3 catalysts, the $$ \Delta G_{\rm{N_{2}}\ast} $$ value is close to the zero, ranging from 0.35 to -5.12 eV, implying that the N2 molecule exhibits favorable adsorption characteristics on these catalysts. In comparison to the original N≡N triple bond length in N2 molecules [Table 1], regardless of whether the N2 molecule is adsorbed in an end-on or side-on configuration, a noticeable elongation is observed in the N≡N triple bond, indicating that the adsorption of N2 on the selected 21 catalysts can attenuate the N≡N, facilitating the activation of the subsequent NRR process. According to Bader charge analysis, Figure 3D and E shows that the TM atoms and N atoms are indicated to undergo a significant charge transfer, and the metal trimer cluster loses an electron ranging from 0.24 to 1.97 e [Table 1]. As shown in Figure 3F, the adsorption energy of N2 is linearly related to the d-orbital center (εd) of the metal clusters. The higher the d orbital center of the TM3@C3N3 system, the more favorable its interaction with the π* orbital of N2 and provides more charge transfer to the N2 molecule, thus showing a stronger binding capacity for N2 adsorption[69-72]. Guided by these essential parameters, the catalytic response of TM trimer clusters can be finely regulated and manipulated by designing moderate cluster-carrier interactions through the selection of suitable metal elements.

Reaction mechanism of electrochemical NRR

As shown in Figure 4A, one reaction mechanism for the electrocatalytic synthesis of ammonia, namely the enzymatic mechanism, is depicted, and an additional consecutive mechanism is presented in Supplementary Figure 1A. The distal mechanism and the alternative mechanism are also detailed in Supplementary Figure 1B. A common feature among them is that the N atom at one end of N2 is adsorbed on the catalyst, while the N element at the other end is not adsorbed. However, the N element at the far end preferentially reacts with the H proton. In the distal mechanism, N atoms leaving the surface of the catalyst preferentially react with H protons and release as ammonia, leaving *N adsorbed on the catalyst, which will start hydrogenation and form ammonia. An alternative mechanism involves hydrogenating two N atoms using six proton-electron pairs to form two NH3 molecules. Contrarily, in the enzymatic pathway, the N2 molecule is first decomposed into adsorbed *N atoms, and then, *N is gradually hydrogenated into ammonia. The hydrogenation process follows the same pathway as the alternative mechanism. In the consecutive mechanism, one of the N atoms undergoes hydrogenation first and subsequently reacts with the remaining N atoms, resulting in the formation of the second NH3 molecule. Owing to the distinct structures and performances of catalysts, the NRR reaction mechanism often varies. Nevertheless, numerous studies have shown that the initial or final step can be considered the potential determinant independent of NRR mechanisms[73-75]. The proton step with the maximal positive free energy change (∆Gmax) is defined as the PDS. It is well known that the ideal catalyst for an electrochemical NRR should meet the following criteria: (1) The ∆Gmax of the two key steps is below 0.55 eV, that is, $$ \Delta G_{\ast \rm{N_{2}}\rightarrow \ast \text{NNH}} $$ ≤ 0.55 eV and $$ \Delta G_{\ast \rm{NH_{2}}\rightarrow \ast \rm{NH_{3}}} $$ ≤ 0.55 eV, so the performance of the catalyst in question is superior to that of the best pure metal or nanometal cluster catalyst; (2) $$ E_{\rm ads(\ast N_{2})} - E_{\rm ads(\ast H)} $$ < 0, which proves that N2 has good selectivity in catalysts.

Computational design of spatially confined triatomic catalysts for nitrogen reduction reaction

Figure 4. (A) Schematic illustration of Enzymatic mechanism towards NH3 formation on TM3@C3N3. Diagrams for N2 electroreduction via enzymatic mechanism on (B) Pt3@C3N3 and (C) Ru3@C3N3; (D) The optimized geometry of various intermediates on the Pt3@C3N3 structure along the enzymatic pathway of NRR. NRR: Nitrogen reduction reactions.

For a prompt evaluation of TM3@C3N3 catalysts regarding their NRR performance, we calculated the ∆G for the initial and final stages of NRR under open-circuit conditions (U = 0). The resulting values, along with corresponding diagrams, are presented in Table 2 and Supplementary Figure 2. Four classifications are considered to evaluate the critical steps, with the critical point set at 0.55 eV. Consequently, six systems met the above criteria, demonstrating decent catalytic activity in electrochemical NRR. In the case of Pt3@C3N3, the associated free energy diagram is presented in Figure 4B. The Pt3@C3N3 exhibits a propensity for NRR via enzymatic mechanisms, wherein N2 adopts a side-on configuration. In the enzymatic mechanism, the initial step of hydrogenation involves the formation of a chemical bond between a hydrogen atom and one of the N atoms, followed by alternating bonding to the two N atoms until a second NH3 molecule is produced. The ∆Gmax change is 0.24 eV, establishing the final hydrogenation step as the PDS of the entire electrochemical NRR. Similarly, we investigated the NRR pathway on Ru3@C3N3, as depicted in Figure 4C, revealing a ∆Gmax of 0.35 eV.

Table 2

DFT-calculated potential-determining steps (including reaction mechanisms and corresponding reactions) and corresponding limiting potentials (UL) for TM3@C3N3 (TM = Re, Pt, Ru, Rh, Ta, Ir)

TM3@C3N3Potential-determining stepUL (V)
MechanismReaction
ReConsecutive*HNNH + H+ + e- → *HNNH2-0.11
PtEnzymatic*NH2 + H+ + e- → *NH3-0.24
RuEnzymatic
Consecutive
*N2 + H+ + e- → *NNH
*NNH+ H+ + e- → *NNH2
-0.35
-0.40
RhEnzymatic
Consecutive
*N2 + H+ + e- → *NNH-0.40
TaConsecutive*HNNH + H+ + e- → *HNNH2-0.42
IrEnzymatic
Consecutive
*NH2 + H+ + e- → *NH3-0.54

The relevant intermediate geometries of Pt3@C3N3 for each step are presented clearly in Figure 4D. It is evident that throughout the entire NRR process, the initial step (*N2 → *NNH) is identified as the PDS. Briefly speaking, for Pt3@C3N3 catalysts, the corresponding limiting potential (UL) value is -0.24 V, while for Ru3@C3N3 catalysts, the UL value is -0.35 V. Consequently, by applying the UL to the surfaces of Pt3@C3N3 and Ru3@C3N3 catalysts, it is ensured that all electron transfer steps occur spontaneously without any uphill energy barriers, which is beneficial to the production of NH3, where the reaction process of Ru3@C3N3 also follows an enzymatic mechanism.

It is worth emphasizing that the development of efficient NRR catalysts remains challenging due to the competition with HER[66,76,77]. An ideal NRR catalyst would demonstrate significantly higher NRR activity and considerably lower HER activity. To assess selectivity, we calculated the N2 adsorption energy and hydrogen adsorption energy on the designed TM3@C3N3 catalysts using Equation (2). A more negative difference between N2 adsorption energy and hydrogen adsorption energy indicates higher selectivity for NRR. The results are presented in Supplementary Figure 3. It is evident that both Pt3@C3N3 and Ru3@C3N3, which were previously identified as having the highest NRR activity, exhibit markedly higher selectivity for NRR over HER. These findings imply that these catalysts can ensure a high Faraday efficiency in catalytic electrochemical NRR.

CONCLUSIONS

In summary, this study provides a systematic investigation into the potential of C3N3-loaded TACs in electrocatalytic NRR using DFT calculations. Employing a stringent “five-step” filtering strategy, we have identified TM3@C3N3 (TM = Pt, Ru, Re) as highly promising candidates with the attributes of low energy cost, high selectivity, remarkable stability (both thermodynamic and kinetic), and remarkably low limiting potentials (-0.35~-0.11 V). Our analysis of electronic properties highlights that the exceptional NRR activity can be ascribed to the electron acceptance and donation mechanism involving d-π* interactions. This mechanism, combined with charge density differences and PDOS, underscores the qualifications of TM3@C3N3 as electrocatalysts. Furthermore, we have explored the linkage between chemical activity and the electronic structure of TM3@C3N3 surfaces, revealing a pivotal physical parameter that allows precise control over the catalytic performance of TM trimer clusters. In conclusion, this work significantly improves our overall comprehension regarding the stability, activity, and selectivity of TM3@C3N3 electrocatalysts. Moreover, it provides a valuable methodology for efficiently screening and designing innovative TACs specifically tailored for NRR. We anticipate that these findings will ignite further exploration and experimentation, both in theory and practice, aiming to unlock the vast potential of TACs in NRR and its associated electrochemical reactions.

DECLARATIONS

Authors’ contributions

Made substantial contributions to the conception and design of this article, writing, and editing: Pei W, Zhang W, Yu X

Made substantial contributions to the collation of literature and writing: Liu Y, Zhou S, Tu Y, Zhao J

Performed data analysis and discussion: Hou L, Xia W, Wang Z

Availability of data and materials

Supplementary Materials are available from the Journal of Materials Informatics or from the authors.

Financial support and sponsorship

This work was supported by the National Natural Science Foundation of China (Nos. 12304300, 11974068, 91961204, and 12075201), the Natural Science Foundation of Jiangsu Province (Nos. BK20230555 and BK20230563), the Natural Science Foundation of Jiangsu Higher Education Institutions of China (No. 23KJB140019), the Fundamental Research Funds for the Central Universities of China (Nos. DUT20LAB110 and DUT22ZD103), Outstanding Doctor Program of Yangzhou City “Lv Yang Jin Feng” (No. YZLYJFJH2022YXBSO84), and Open Research Fund of CNMGE Platform & NSCC-TJ (CNMGE2023007). The authors acknowledge the computer resources provided by the Shanghai Supercomputer Center.

Conflicts of interest

All authors declared that there are no conflicts of interest.

Ethical approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Copyright

© The Author(s) 2023.

Supplementary Materials

REFERENCES

1. Galloway JN, Townsend AR, Erisman JW, et al. Transformation of the nitrogen cycle: recent trends, questions, and potential solutions. Science 2008;320:889-92.

2. Klerke A, Christensen CH, Nørskov JK, Vegge T. Ammonia for hydrogen storage: challenges and opportunities. J Mater Chem 2008;18:2304-10.

3. Guo J, Chen P. Catalyst: NH3 as an energy carrier. Chem 2017;3:709-12.

4. Gao S, Liu X, Wang Z, et al. Spin regulation for efficient electrocatalytic N2 reduction over diatomic Fe-Mo catalyst. J Colloid Interface Sci 2023;630:215-23.

5. Zhang G, Zhang X, Meng Y, Pan G, Ni Z, Xia S. Layered double hydroxides-based photocatalysts and visible-light driven photodegradation of organic pollutants: a review. Chem Eng J 2020;392:123684.

6. Chen JG, Crooks RM, Seefeldt LC, et al. Beyond fossil fuel-driven nitrogen transformations. Science 2018;360:eaar6611.

7. Yang X, Nash J, Anibal J, et al. Mechanistic insights into electrochemical nitrogen reduction reaction on vanadium nitride nanoparticles. J Am Chem Soc 2018;140:13387-91.

8. Gao Y, Zhuo H, Cao Y, et al. A theoretical study of electrocatalytic ammonia synthesis on single metal atom/MXene. Chinese J Catal 2019;40:152-9.

9. Wu J, Li JH, Yu YX. Single Nb or W atom-embedded BP monolayers as highly selective and stable electrocatalysts for nitrogen fixation with low-onset potentials. ACS Appl Mater Interfaces 2021;13:10026-36.

10. Du P, Huang Y, Zhu G, et al. Nitrogen reduction reaction on single cluster catalysts of defective PC6-trimeric or tetrameric transition metal. Phys Chem Chem Phys 2022;24:2219-26.

11. Andersen SZ, Čolić V, Yang S, et al. A rigorous electrochemical ammonia synthesis protocol with quantitative isotope measurements. Nature 2019;570:504-8.

12. Zhang L, Meng Y, Shen H, et al. High-efficiency photocatalytic ammonia synthesis by facet orientation-supported heterojunction Cu2O@BiOCl[100] boosted by double built-in electric fields. Inorg Chem 2022;61:6045-55.

13. Li L, Tang C, Yao D, Zheng Y, Qiao SZ. Electrochemical nitrogen reduction: identification and elimination of contamination in electrolyte. ACS Energy Lett 2019;4:2111-6.

14. Hao YC, Guo Y, Chen LW, et al. Promoting nitrogen electroreduction to ammonia with bismuth nanocrystals and potassium cations in water. Nat Catal 2019;2:448-56.

15. Singh AR, Rohr BA, Statt MJ, Schwalbe JA, Cargnello M, Nørskov JK. Strategies toward selective electrochemical ammonia synthesis. ACS Catal 2019;9:8316-24.

16. Martín AJ, Shinagawa T, Pérez-ramírez J. Electrocatalytic reduction of nitrogen: from haber-bosch to ammonia artificial leaf. Chem 2019;5:263-83.

17. Han B, Meng H, Li F, Zhao J. Fe3 cluster anchored on the C2N monolayer for efficient electrochemical nitrogen fixation. Catalysts 2020;10:974.

18. Yu L, Li F. Pt2 dimer anchored vertically in defective BN monolayer as an efficient catalyst for N2 reduction: a DFT study. Catalysts 2022;12:1387.

19. Geng Z, Liu Y, Kong X, et al. Achieving a record-high yield rate of 120.9 $$ \mu \mathrm{g}_{\mathrm{NH}_{3}} \cdot \mathrm{mg}_{\text {cat.}}{}^{-1} \cdot \mathrm{h}^{-1} $$ for N2 electrochemical reduction over Ru single-atom catalysts. Adv Mater 2018;30:1803498.

20. Tao H, Choi C, Ding LX, et al. Nitrogen fixation by Ru single-atom electrocatalytic reduction. Chem 2019;5:204-14.

21. Han Z, Huang S, Zhang J, et al. Single Ru-N4 site-embedded porous carbons for electrocatalytic nitrogen reduction. ACS Appl Mater Interfaces 2023;15:13025-32.

22. Nørskov JK, Bligaard T, Hvolbaek B, Abild-Pedersen F, Chorkendorff I, Christensen CH. The nature of the active site in heterogeneous metal catalysis. Chem Soc Rev 2008;37:2163-71.

23. Azofra LM, Li N, MacFarlane DR, Sun C. Promising prospects for 2D d2-d4 M3C2 transition metal carbides (MXenes) in N2 capture and conversion into ammonia. Energy Environ Sci 2016;9:2545-9.

24. Choi C, Back S, Kim NY, Lim J, Kim YH, Jung Y. Suppression of hydrogen evolution reaction in electrochemical N2 reduction using single-atom catalysts: a computational guideline. ACS Catal 2018;8:7517-25.

25. Skúlason E, Bligaard T, Gudmundsdóttir S, et al. A theoretical evaluation of possible transition metal electro-catalysts for N2 reduction. Phys Chem Chem Phys 2012;14:1235-45.

26. Liu L, Corma A. Metal catalysts for heterogeneous catalysis: from single atoms to nanoclusters and nanoparticles. Chem Rev 2018;118:4981-5079.

27. Su P, Pei W, Wang X, et al. Exceptional electrochemical HER performance with enhanced electron transfer between Ru nanoparticles and single atoms dispersed on a carbon substrate. Angew Chem Int Ed Engl 2021;60:16044-50.

28. Yang XF, Wang A, Qiao B, Li J, Liu J, Zhang T. Single-atom catalysts: a new frontier in heterogeneous catalysis. Acc Chem Res 2013;46:1740-8.

29. Lv X, Wei W, Wang H, Huang B, Dai Y. Holey graphitic carbon nitride (g-CN) supported bifunctional single atom electrocatalysts for highly efficient overall water splitting. Appl Catal B Environ 2020;264:118521.

30. Zhao J, Chen Z. Single Mo atom supported on defective boron nitride monolayer as an efficient electrocatalyst for nitrogen fixation: a computational study. J Am Chem Soc 2017;139:12480-7.

31. Zhao J, Zhao J, Cai Q. Single transition metal atom embedded into a MoS2 nanosheet as a promising catalyst for electrochemical ammonia synthesis. Phys Chem Chem Phys 2018;20:9248-55.

32. Chen Z, Zhao J, Cabrera CR, Chen Z. Computational screening of efficient single-atom catalysts based on graphitic carbon nitride (g-C3N4) for nitrogen electroreduction. Small Methods 2019;3:1800368.

33. He T, Matta SK, Will G, Du A. Transition-metal single atoms anchored on graphdiyne as high-efficiency electrocatalysts for water splitting and oxygen reduction. Small Methods 2019;3:1800419.

34. Liu Y, Xu Q, Fan X, et al. Electrochemical reduction of N2 to ammonia on Co single atom embedded N-doped porous carbon under ambient conditions. J Mater Chem A 2019;7:26358-63.

35. Wang S, Wei W, Lv X, Huang B, Dai Y. W supported on g-CN manifests high activity and selectivity for N2 electroreduction to NH3. J Mater Chem A 2020;8:1378-85.

36. Yu X, Pei W, Xu W, Zhao Y, Su Y, Zhao J. Core-packing-related vibrational properties of thiol-protected gold nanoclusters and their excited-state behavior. Inorg Chem 2023;62:20450-7.

37. Zhao J, Du Q, Zhou S, Kumar V. Endohedrally doped cage clusters. Chem Rev 2020;120:9021-163.

38. Yu X, Sun Y, Xu W, et al. Tuning photoelectron dynamic behavior of thiolate-protected MAu24 nanoclusters via heteroatom substitution. Nanoscale Horiz 2022;7:1192-200.

39. Yan H, Lin Y, Wu H, et al. Bottom-up precise synthesis of stable platinum dimers on graphene. Nat Commun 2017;8:1070.

40. Li G, Li Y, Liu H, Guo Y, Li Y, Zhu D. Architecture of graphdiyne nanoscale films. Chem Commun 2010;46:3256-8.

41. Tian S, Fu Q, Chen W, et al. Carbon nitride supported Fe2 cluster catalysts with superior performance for alkene epoxidation. Nat Commun 2018;9:2353.

42. Ji S, Chen Y, Fu Q, et al. Confined pyrolysis within metal-organic frameworks to form uniform Ru3 clusters for efficient oxidation of alcohols. J Am Chem Soc 2017;139:9795-8.

43. Pei W, Hou L, Yu X, et al. Graphitic carbon nitride supported trimeric metal clusters as electrocatalysts for N2 reduction reaction. J Catal 2024;429:115232.

44. Li Y, Zhang Q, Li C, et al. Atomically dispersed metal dimer species with selective catalytic activity for nitrogen electrochemical reduction. J Mater Chem A 2019;7:22242-7.

45. Li H, Zhao Z, Cai Q, Yin L, Zhao J. Nitrogen electroreduction performance of transition metal dimers embedded into N-doped graphene: a theoretical prediction. J Mater Chem A 2020;8:4533-43.

46. Zheng G, Li L, Tian Z, Zhang X, Chen L. Heterogeneous single-cluster catalysts (Mn3, Fe3, Co3, and Mo3) supported on nitrogen-doped graphene for robust electrochemical nitrogen reduction. J Energy Chem 2021;54:612-9.

47. Cui C, Zhang H, Luo Z. Nitrogen reduction reaction on small iron clusters supported by N-doped graphene: a theoretical study of the atomically precise active-site mechanism. Nano Res 2020;13:2280-8.

48. Li L, Xu L. Design of a graphene nitrene two-dimensional catalyst providing a well-defined site accommodating up to three metals, with application to N2 reduction electrocatalysis. Chem Commun 2020;56:8960-3.

49. Kresse G, Hafner J. Ab initio molecular dynamics for liquid metals. Phys Rev B Condens Matter 1993;47:558-61.

50. Kresse G, Hafner J. Ab initio molecular-dynamics simulation of the liquid-metal-amorphous-semiconductor transition in germanium. Phys Rev B Condens Matter 1994;49:14251-69.

51. Kresse G, Joubert D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys Rev B 1999;59:1758.

52. Blöchl PE. Projector augmented-wave method. Phys Rev B Condens Matter 1994;50:17953-79.

53. Perdew JP, Burke K, Ernzerhof M. Generalized gradient approximation made simple. Phys Rev Lett 1996;77:3865.

54. Tian J, Hou L, Xia W, et al. Solar driven CO2 hydrogenation to HCOOH on (TiO2)n (n = 1-6) atomic clusters. Phys Chem Chem Phys 2023;25:28533-40.

55. Grimme S, Antony J, Ehrlich S, Krieg H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J Chem Phys 2010;132:154104.

56. Grimme S, Ehrlich S, Goerigk L. Effect of the damping function in dispersion corrected density functional theory. J Comput Chem 2011;32:1456-65.

57. Monkhorst HJ, Pack JD. Special points for Brillouin-zone integrations. Phys Rev B 1976;13:5188.

58. Peterson AA, Abild-Pedersen F, Studt F, Rossmeisl J, Nørskov JK. How copper catalyzes the electroreduction of carbon dioxide into hydrocarbon fuels. Energy Environ Sci 2010;3:1311-5.

59. Peterson AA, Nørskov JK. Activity descriptors for CO2 electroreduction to methane on transition-metal catalysts. J Phys Chem Lett 2012;3:251-8.

60. Nørskov JK, Rossmeisl J, Logadottir A, et al. Origin of the overpotential for oxygen reduction at a fuel-cell cathode. J Phys Chem B 2004;108:17886-92.

61. Lv X, Wei W, Huang B, Dai Y, Frauenheim T. High-throughput screening of synergistic transition metal dual-atom catalysts for efficient nitrogen fixation. Nano Lett 2021;21:1871-8.

62. Zhou Y, Wang J, Liang L, et al. Unraveling the size-dependent effect of Ru-based catalysts on ammonia synthesis at mild conditions. J Catal 2021;404:501-11.

63. Wang S, Shi L, Bai X, Li Q, Ling C, Wang J. Highly efficient photo-/electrocatalytic reduction of nitrogen into ammonia by dual-metal sites. ACS Cent Sci 2020;6:1762-71.

64. Ling C, Niu X, Li Q, Du A, Wang J. Metal-free single atom catalyst for N2 fixation driven by visible light. J Am Chem Soc 2018;140:14161-8.

65. Hu R, Li Y, Zeng Q, Wang F, Shang J. Bimetallic pairs supported on graphene as efficient electrocatalysts for nitrogen fixation: search for the optimal coordination atoms. ChemSusChem 2020;13:3636-44.

66. Pei W, Zhou S, Zhao J, Du Y, Dou SX. Optimization of photocarrier dynamics and activity in phosphorene with intrinsic defects for nitrogen fixation. J Mater Chem A 2020;8:20570-80.

67. Janesko BG. Replacing hybrid density functional theory: motivation and recent advances. Chem Soc Rev 2021;50:8470-95.

68. Shinde R, Yamijala SSRKC, Wong BM. Improved band gaps and structural properties from Wannier-Fermi-Löwdin self-interaction corrections for periodic systems. J Phys Condens Matter 2021;33:115501.

69. Back S, Lim J, Kim NY, Kim YH, Jung Y. Single-atom catalysts for CO2 electroreduction with significant activity and selectivity improvements. Chem Sci 2017;8:1090-6.

70. Li H, Pei W, Yang X, Zhou S, Zhao J. Pt overlayer for direct oxidation of CH4 to CH3OH. Chinese Chem Lett 2023;34:108292.

71. Zhou S, Pei W, Du Q, Zhao J. Foreign atom encapsulated Au12 golden cages for catalysis of CO oxidation. Phys Chem Chem Phys 2019;21:10587-93.

72. Pei W, She J, Yu X, Zhou S, Zhao J. Atomically precise gold nanoclusters for CO oxidation: balancing activity and stability by ligand shedding. J Phys D Appl Phys 2023;56:445304.

73. Han B, Li F. Regulating the electrocatalytic performance for nitrogen reduction reaction by tuning the N contents in Fe3@NxC20-x (x = 0~4): a DFT exploration. J Mater Inf 2023;3:24.

74. Gu YT, Gu YM, Tao Q, Wang X, Zhu Q, Ma J. Machine learning for prediction of CO2/N2/H2O selective adsorption and separation in metal-zeolites. J Mater Inf 2023;3:19.

75. Hu Y, Chen J, Wei Z, He Q, Zhao Y. Recent advances and applications of machine learning in electrocatalysis. J Mater Inf 2023;3:18.

76. Sun Y, Pei W, Xie M, et al. Excitonic Au4Ru2(PPh3)2(SC2H4Ph)8 cluster for light-driven dinitrogen fixation. Chem Sci 2020;11:2440-7.

77. Zhou S, Pei W, Zhao Y, Yang X, Liu N, Zhao J. Low-dimensional non-metal catalysts: principles for regulating p-orbital-dominated reactivity. npj Comput Mater 2021;7:186.

Cite This Article

Export citation file: BibTeX | RIS

OAE Style

Pei W, Zhang W, Yu X, Hou L, Xia W, Wang Z, Liu Y, Zhou S, Tu Y, Zhao J. Computational design of spatially confined triatomic catalysts for nitrogen reduction reaction. J Mater Inf 2023;3:26. http://dx.doi.org/10.20517/jmi.2023.35

AMA Style

Pei W, Zhang W, Yu X, Hou L, Xia W, Wang Z, Liu Y, Zhou S, Tu Y, Zhao J. Computational design of spatially confined triatomic catalysts for nitrogen reduction reaction. Journal of Materials Informatics. 2023; 3(4): 26. http://dx.doi.org/10.20517/jmi.2023.35

Chicago/Turabian Style

Pei, Wei, Wenya Zhang, Xueke Yu, Lei Hou, Weizhi Xia, Zi Wang, Yongfeng Liu, Si Zhou, Yusong Tu, Jijun Zhao. 2023. "Computational design of spatially confined triatomic catalysts for nitrogen reduction reaction" Journal of Materials Informatics. 3, no.4: 26. http://dx.doi.org/10.20517/jmi.2023.35

ACS Style

Pei, W.; Zhang W.; Yu X.; Hou L.; Xia W.; Wang Z.; Liu Y.; Zhou S.; Tu Y.; Zhao J. Computational design of spatially confined triatomic catalysts for nitrogen reduction reaction. J. Mater. Inf. 2023, 3, 26. http://dx.doi.org/10.20517/jmi.2023.35

About This Article

Special Issue

This article belongs to the Special Issue Theory-Assisted Developing Electrocatalysis
© The Author(s) 2023. Open Access This article is licensed under a Creative Commons Attribution 4.0 International License (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, sharing, adaptation, distribution and reproduction in any medium or format, for any purpose, even commercially, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Data & Comments

Data

Views
452
Downloads
169
Citations
0
Comments
0
6

Comments

Comments must be written in English. Spam, offensive content, impersonation, and private information will not be permitted. If any comment is reported and identified as inappropriate content by OAE staff, the comment will be removed without notice. If you have any queries or need any help, please contact us at support@oaepublish.com.

0
Download PDF
Cite This Article 5 clicks
Like This Article 6 likes
Share This Article
Scan the QR code for reading!
See Updates
Contents
Figures
Related
Journal of Materials Informatics
ISSN 2770-372X (Online)
Follow Us

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/