Open Access Article
Kiruthika
Pandiyan
and
Hsin-Tsung
Chen
*
Department of Chemistry, Research Center for Semiconductor Materials and Advanced Optics, and R&D Center for Membrane Technology, Chung Yuan Christian University, Chungli District, Taoyuan City 320314, Taiwan. E-mail: htchen@cycu.edu.tw; Tel: +886-3-265-3324
First published on 20th November 2025
The electrochemical nitrate reduction reaction (NO3RR) offers a sustainable approach for converting nitrate pollutants into valuable ammonia under ambient conditions. Herein, we employ density functional theory (DFT) to systematically investigate the catalytic potential of homonuclear transition metal dual-atom catalysts (DACs) anchored on biphenylene (TM2@BPN) for NO3RR. Among 28 candidates, five DACs: Mo2@BPN, Ru2@BPN, Rh2@BPN, Os2@BPN, and Ir2@BPN, exhibit low limiting potentials (−0.40 to −0.16 V) and exceptional ammonia selectivity. Rh2@BPN, in particular, achieves a theoretical faradaic efficiency of 100%, effectively suppressing competing hydrogen evolution. Electronic analyses reveal that dual-site π-donation/π*-back-donation interactions, d-band center tuning, and charge redistribution collectively enhance NO3− activation compared to single-atom analogues. Importantly, descriptor-based volcano plots (ΔG*NO3, εd, and ψ) establish generalizable design rules that correlate electronic structure with catalytic trends, enabling predictive DAC screening beyond exhaustive pathway calculations. Ab initio molecular dynamics simulations further confirm the thermal stability of the most active DACs. This work introduces biphenylene as a robust support for stabilizing DACs, establishes structure–property–performance correlations, and provides transferable mechanistic and descriptor-based insights for the rational design of next-generation catalysts for selective multi-electron electrochemical transformations such as nitrate-to-ammonia conversion.
Dual-atom catalysts (DACs), comprising two adjacent metal atoms, have recently emerged as a promising class of materials that combine the advantages of SACs with synergistic effects. These dual sites enable independent optimization of multiple reaction steps, making DACs particularly attractive for complex processes like NO3RR.24–28 For example, the Cu2 embedded on n-doped graphene was experimentally verified and exhibited good activity for NO3RR with 97.4% faradaic efficiency.29 Fe/Cu diatomic catalysts on nitrogen-doped graphene achieving faradaic efficiencies up to 92.5% at −0.3 V vs. RHE demonstrate strong potential, yet they often rely on heterogeneous dual-atom sites and systematic prediction and design of homonuclear DACs has been less explored.30 Additionally, high-throughput computational screening of M1M2@g-CN DACs (eg FeMo@g-CN, CrMo@g-CN) has identified promising heterogeneous dual-metal combinations, but descriptors correlating electronic structure to performance are still diverse.31 Moreover, reviews of NO3RR underscore that mechanical understanding, especially linking the electronic structure of active dual metal sites, to ammonia selectivity, remains underdeveloped.32 Despite recent progress, rational screening of DACs for NO3RR remains limited, and suitable supports that can stabilize and electronically engage dual-metal centers are critically needed. In particular, the potential of homonuclear DACs for NO3RR remains largely unexplored, how the electronic structure of homonuclear DACs influences the activity and selectivity towards ammonia formation, and reliable electronic descriptors for predicting their catalytic behavior are still lacking.
Herein, we explore biphenylene (BPN)—a recently synthesized sp2-hybridized carbon allotrope featuring four-, six-, and eight-membered rings—as a novel support for homonuclear transition metal DACs. Unlike graphene and other 2D carbon materials the intrinsic ring topology provides a porous lattice with a large number of low-coordination sites that can efficiently stabilize metal atoms. In specific cases, BPN outperforms graphene in catalytic performance due to its higher σ-center position (−7.49 eV vs. −8.09 eV), which allows for greater interaction with reaction intermediates. BPN offers high thermal stability with high thermal conductivity and electrical stability and unique electronic properties including metallic behavior with a tilted Dirac cone above the Fermi level and strong in-plane anisotropy conducive to electrocatalysis.33–37 BPN is a promising catalyst for various reactions, including HER,38 OER,39 CO2RR,40 and NRR.41 In addition, BPN can anchor transition-metal atoms without needing engineered defects or dopant heteroatoms, leading to highly active single-atom Pd@Bip catalysts with very low overpotentials for HER (∼0.05 V) and OER (∼0.50 V), outperforming many conventional catalysts.42 Given the unique advantages of BPN, it is desirable to investigate the potential of BPN for NO3RR and its underlying catalytic mechanism, which has not been reported until now.
Using density functional theory (DFT), we systematically investigate 28 TM2@BPN catalysts for NO3RR, evaluating their activity, selectivity, electronic structure, and thermal stability. After assessing 28 candidates of TM2@BPN, 21 homonuclear DACs that exhibit thermodynamic stability were selected for further investigation. Our screening identifies five promising DACs, Mo2@BPN, Ru2@BPN, Rh2@BPN, Os2@BPN, and Ir2@BPN that exhibit low limiting potentials and excellent selectivity toward NH3. In particular, Rh2@BPN demonstrates a theoretical faradaic efficiency of 100%. Electronic structure analysis, descriptor-based volcano plots, and ab initio molecular dynamics simulations reveal structure–property–performance relationships that provide fundamental insight and design principles for next-generation NO3RR catalysts.
| Eads = Etotal − Esubstrate − Eadsorbate | (1) |
| ΔG = ΔE + ΔEZPE − TΔS + ΔGU + ΔGpH | (2) |
ΔGU is the applied potential with 0 V in this study, and ΔGpH is considered as the correction free energy of H+, which can be calculated using the equation that follows:
| ΔGpH = 2.303 × kBT × pH | (3) |
| UL = −ΔGmax/e | (4) |
| ΔG*NO3 = G*NO3 − G* − GHNO3 (g) + 1/2GH2 (g) + ΔGcorrect | (5) |
![]() | ||
| Fig. 1 (a) Optimized structure of BPN substrate, with C1 and C2 carbon highlighted. (b) The electron localization function (ELF) of BPN. | ||
As shown in Fig. 1b, the electron localization function of BPN ranges from 0 au (blue) to 1 au (red), where higher values indicate stronger electron localization. The BPN showed discrete regions with values of roughly 0.5 au, indicating delocalized electrical structure. The C–C bond has a maximum value of around 0.9 au, indicating the presence of a strong covalent connection in BPN.
The TM2@BPN system was designed by anchoring 28 TM dimers (TM = 3d, 4d, 5d) to the BPN surface. As shown in Fig. S1, four potential TM dimer anchoring configurations were considered: (I) C1–C1 bridging in a six-membered ring, (II) C1–C1 bridging in an eight-membered ring, (III) top C1 site in an eight-membered ring, and (IV) C2–C2 bridging in an eight-membered ring. The stability and relative energy of these TM dimer anchoring configurations have been thoroughly investigated in previous reports.41Fig. 2a depicts the most stable anchoring configurations after the structural optimization. Using the most energetically favorable configurations, we investigated the stability of TM2@BPN by calculating the binding and cohesive energy. To assess the binding strength between the TM dimer and the BPN substrate, the binding energy (Ebin) was calculated as follows:
| Ebin = ETM2@BPN − EBPN − 2ETM | (6) |
| Ecoh = (Ebulk − nETM)/n | (7) |
![]() | ||
| Fig. 2 (a) The stable binding configurations of TM2@BPN, with 28 transition metals considered for anchoring on BPN substrate, (b) Binding and cohesive energy comparison plot for TM2@BPN. | ||
![]() | ||
| Fig. 3 (a) Comparison of adsorption free energies of NO3−, and H on TM2@BPN. (b) The scaling relationship between NO3− adsorption free energies and charge transfer of TM dimers. | ||
![]() | ||
| Fig. 4 Detailed reaction pathway for NH3 production, as well as NO dimer pathway for N2O, and N2 production. | ||
Descriptor based catalyst design serves as a more effective and focused approach for electrocatalytic development and screening.55 The Gibbs free energies of intermediates, which play a critical role in chemical reactions, can function as a primary indicator for predicting the catalytic activity. Therefore, the Gibbs free energy of NO3− is chosen as a descriptor, as nitrate is a key intermediate and its adsorption is an essential step. Fig. 6 shows the graph of limiting potential plotted as a function of the Gibbs free energy of NO3−. Apparently, there is a distinct volcano shaped relationship between them with Rh2@BPN precisely situated at the top of the volcano plot. Apart from, Rh2@BPN additional TM2@BPN systems (highlighted in purple in Fig. 6) are characterized by low limiting potentials (UL). We can deduce that, insufficient adsorption of the catalyst results in a potential determining step (PDS) from *NO to *NOH, while intense adsorption results in high energy demanding step from *NH2 to *NH3/*OH to *OH2. This analysis is consistent with the Sabatier principle, which states that the ideal catalytic activity is achieved if the essential intermediates and catalysts bind with moderate intensity. Accordingly, catalysts with moderate adsorption (ΔG*NO3 between −1.5 to 0.9 eV) have high catalytic activity.
![]() | ||
| Fig. 6 NO3RR Volcano plot of TM2@BPN with the descriptor of Gibbs free energy NO3 (ΔG*NO3) vs. limiting potential (UL). | ||
We determined the Gibbs free energy profiles for NO3RR on Rh2@BPN, Os2@BPN, and Ir2@BPN, (Fig. 7a–c) surfaces in order to gain more insight into the reaction mechanism of nitrate reduction to NH3. The DFT-optimized configurations of intermediates associated with the free energy diagram of Rh2@BPN, Os2@BPN and Ir2@BPN are shown in Fig. S3, the computed zero-point energy and entropy are listed in Table S9–S11, and the Gibbs free energy profiles for the remaining DACs are shown in Fig. S4–S9. It is evident that NO3− is strongly chemisorbed on Rh2@BPN, Os2@BPN, and Ir2@BPN via two TM–O bonds, with free energy changes of −1.52, −1.21, −1.04 eV respectively. On Rh2@BPN, the initial proton–electron coupling preferentially attacks the nitrate oxygen interacting with Rh over terminal oxygen. As a result, NO3 dissociates to form *OH–*NO2 with an energy release of −1.35 eV. A similar trend can be identified on Os2@BPN and Ir2@BPN with an exothermic energy change of −2.88 eV, and −2.09 eV respectively. The second hydrogenation step leads to the formation of H2O, which possesses greater energetic stability than the hydrogenation of the *OH–*NO2 molecule to *OH–*NO2H. This step is exothermic by −1.19 eV for Rh2@BPN, and leaves the *NO2 on the active site with two Rh–O bonds. Then *NO2 undergoes hydrogenation in the third step, analogous to that of *NO3− to form *OH–*NO with an energy release of −1.40 eV. The fourth hydrogenation results in the formation of *NO on the Rh–Rh bridging site, following the release of the second water molecule. Three different bonding modes were assessed for *NO adsorption: **NO (binding via the NO-side), *ON (binding via the O-end), and *NO (binding via the N-end) as shown in Fig. S10. Depending on the *NO configuration (NO-end, NO-side, and ON-end), the following hydrogenation will adopt a different pathway. The computed Gibbs free energy for all the binding modes of *NO is tabulated in Table S12. The findings indicate that *NO and **NO modes are thermodynamically more favorable than *ON mode. On Rh2@BPN, *NO is more likely to adsorb via side-on mode (**NO). The hydrogenation of **NO requires an energy input of 0.16 eV to generate **HNO on the active site. Proceeding with the reactions, steps six to eight are enzymatic protonation of the intermediates (*NH–*O → *NH–*OH → *NH2–*OH → *NH2), with corresponding energy changes of −0.34, −1.06, and −0.63 eV respectively, to release the third water molecule and leave *NH2 on the active site. Eventually the *NH2 can be protonated to generate *NH3 with a Gibbs free energy change of 0.04 eV. The potential determining step of NO3RR on Rh2@BPN is the conversion of *NO → *HNO with a low UL of −0.16 eV.
On Os2@BPN, the hydrogenation of *OH–*NO2 results in the formation of NO2 on the active site with two Os–O bonds, this step involves an endothermic contribution of 0.26 eV. Further hydrogenation results in the formation of *NO. On Os2@BPN *NO adsorption prefers the end on mode (*NO) with an energy release of −0.53 eV. The hydrogenation of *NO requires an energy input of 0.32 eV to generate *NOH on the active site. The ensuing hydrogenations are all exothermic and proceed along the *NOH → *N → *NH → *NH2 → *NH3 pathway. The potential determining step of NO3RR on Os2@BPN is the conversion of *NO → *HNO with a UL of −0.32 eV. For Ir2@BPN the adsorption configuration of *NO3 and *NO intermediates and reaction steps throughout the entire protonation pathway are consistent with Os2@BPN. During the entire protonation mechanism on Ir2@BPN, the conversion of NH2 → NH3 requires a maximum energy input of 0.35 eV, identifying it as the potential-determining step. We have performed a climbing-image nudged elastic band (CI-NEB)56,57 calculations for the potential-determining step of the most active catalysts (Rh2@BPN and Os2@BPN). The results obtained from calculations, are as summarized in Fig. S11 and S12. The calculated activation barrier for the PDS was found to be nearly zero, suggesting that the reaction proceeds without a significant kinetic barrier. Therefore, the process is considered kinetically facile, and our discussion primarily focuses on its thermodynamic aspects.
Furthermore, we extensively evaluated the catalytic activity of the related SACs (Rh@BPN, OS@BPN, and Ir@BPN). Our investigation began with identifying the stable binding site for SACs on BPN, and we found that the single metal preferentially binds on the four-membered ring rather than other binding sites, which is consistent with previous findings.58 Subsequently, for nitrate adsorption, we considered two possible configurations, in the (1-O) configuration, one oxygen atom binds to the metal, while in the (2-O) configuration, two oxygen atoms bind to the metal. Among which the (2-O) configuration is the most favorable on SACs, as shown in Fig. S13 and Table S13. Therefore, the 2-O configuration was selected for further reaction pathway analysis. The energy profile diagrams are shown in Fig. S14, the PDS of Rh@BPN, Os@BPN, and Ir@BPN are all *NO → *NOH, with the UL of −1.15, −1.28, and −0.98 V, respectively. It is apparent that the performance of DACs exhibits much better catalytic activity compared to their SACs catalysts and most of the reported electrocatalysts (Table S14).
As discussed earlier, the NO3RR can lead to multiple byproducts, therefore it is important to investigate the reaction pathways in more detail. Consequently, we investigated the Rh2@BPN, Os2@BPN, and Ir2@BPN catalyst's selectivity for possible N-containing byproducts, such as NO2, NO, N2O, and N2. Fig. 7 illustrates that the production of NO2 on Rh2@BPN, Os2@BPN and Ir2@BPN, from *OH–*NO2 is not energetically feasible due to a substantial energy input of 1.11, 2.16 and 1.28 eV respectively, while the hydrogenation of *OH–*NO2 is thermodynamically favorable. Additionally, a large energy input of 2.25, 2.05 and 1.80 eV were required to overcome the barrier for the direct desorption of NO2 from Rh2@BPN, Os2@BPN and Ir2@BPN respectively. Then, the removal of NO from *OH–*NO intermediate demands high energy input of 3.09, 1.58, and 1.43 eV on Rh2@BPN, Os2@BPN and Ir2@BPN respectively. Following this, the hydrogenation of *NO on Rh2@BPN, Os2@BPN and Ir2@BPN is slightly endothermic, with free energy changes of 0.16, 0.32, and 0.29 eV, respectively. Applying a small potential enables further hydrogenation, which is much more favorable thermodynamically than the direct desorption of NO which is high, at 2.60, 2.09, and 2.33 eV, resembling those of NO2. Thus, *NO will be further hydrogenated on Rh2@BPN, Os2@BPN and Ir2@BPN. Conversely, the NO dimer (N2O2), produced via NO–NO coupling, serves as a key precursor for the generation of N2O and N2. After optimization, the NO dimer undergoes bond dissociation, causing N–N bond cleavage. This implies that the NO dimer is unlikely to form, therefore N2O and N2 byproduct formation is also infeasible. In summary, the generation of byproducts (NO2, NO, N2O, and N2) is effectively suppressed, indicating the high NH3 selectivity of Rh2@BPN, Os2@BPN and Ir2@BPN.
| Δρ = ρ(NO3–TM2@BPN) − ρ(TM2@BPN) − ρ(NO3) | (8) |
We will now analyze the origin of TM2@BPN activity from the standpoint of orbital interactions. Fig. S15b provides a schematic representation of the nitrate adsorption mechanism on the TM dimer substrate. In transition metal active sites, the combination of occupied and vacant d orbitals accounts for their strong binding to NO3.The vacant d orbital of TM atoms can accept electrons from the bonding orbital of NO3, strengthening the TM–O adsorption, whereas the occupied d-orbitals of TM atoms contribute electrons back into the empty π* orbital of NO3, thereby weakening the NO bond. Thus, DACs can maximize the activation potential by utilizing the d orbital of the TM dimer site through a “pull–pull” effect, or they can use the intrinsic dual sites to increase the adjustment space.59 Consequently, the TM dimer active site can alter the local electrical environment of the NO3RR active site.
Fig. 8a and b presents the partial density of states (PDOS) for NO3− adsorbed on Rh2@BPN, and Os2@BPN catalysts, while Fig. S17 displays the PDOS for the remaining systems. The PDOS analysis indicates strong π-donation from NO3− to the metal center, as evidenced by the overlap between the occupied NO3−-2p orbitals and the vacant TM-d orbitals below the Fermi level. Furthermore, the small overlap above the Fermi level between the vacant NO3−-2p orbital and the filled TM-d orbitals indicates π*-back-donation. The d–π interaction facilitates mutual electron transfer, strengthening the interaction between NO3− and TM atoms and thereby stabilizing the adsorption complex. The adsorption strength is largely determined by the occupation of antibonding states. The projected crystal orbital Hamilton population (pCOHP) analysis (Fig. 8c and d) indicates that the interaction between the transition metal and nitrate exhibits predominant bonding characteristics, with a strong bonding contribution below the Fermi level. The antibonding states are primarily located above the Fermi level, indicating that they are largely unoccupied, this implies a strong metal–adsorbate interaction.
In light of the previous explanation, transition metals' d-orbitals are important for molecular adsorption, and their energy distribution is commonly defined by the d-band center (εd).60 It is widely recognized as an effective descriptor to characterize the interaction strength between small molecules and catalytic sites.55 Therefore, we analyzed the correlation between NO3− free energies and the d-band center to gain insights into NO3− activation (Fig. 9 a). The findings indicate that ΔG*NO3 and εd of the transition metal atoms have a linear connection. When εd approaches the Fermi level, the bonding states are primarily occupied while the antibonding states are largely unoccupied, resulting in greater adsorption. On the other hand, when εd is well below the Fermi level most of the metal's d-orbitals are filled causing the electrons to fill both bonding and antibonding orbitals formed during adsorption, causing weaker adsorption. As a result, Pt with the highest d band center (−4.58 eV), has the lowest adsorption energy (−0.91 eV). Conversely, the metal with the highest adsorption energy Ti (−3.20 eV), has the lowest d-band center (0.57 eV). Rh2@BPN, and Os2@BPN with optimal d-band centers of −2.37 and −2.64 eV respectively, exhibit moderate NO3− adsorption energies, which facilitate efficient NO3RR. Fig. 9b shows a plot of εdvs. UL for TM2@BPN, with Rh2@BPN at the top of the volcano, indicating superior catalytic performance. The volcano plot highlights that the excellent NO3RR performance of DACs results from the optimal positioning of the d-band center. Additionally, we compared the d band center of SACs and DACs in order to examine the mechanism of the synergistic effect between the TM dimer site; the results are provided in Table S15. In the case of Rh@BPN, the d-band center is located at −3.26 eV, which resulted in weak nitrate adsorption (−0.96 eV). However, introducing a second Rh atom (Rh2@BPN) shifts the d-band center closer to the Fermi level, which enhances the adsorption strength. This d-band shift significantly affects the adsorption energies of the reaction intermediates, thereby altering the energy barrier of the potential determining step.
As we have discussed earlier, the d-band is crucial for controlling catalytic activity since its occupancy affects the strength of intermediate adsorption, which in turn regulates the overall efficacy of the reaction. The quantity of d-orbital electrons (Nd), influence the adsorption strength, more d-electrons lower the d-band center, there by weakening the adsorbate binding, whereas less d-electrons enhance it, strengthening adsorption. However, Nd alone cannot explain the activity trend, this effect is further modulated by electronegativity (ETM), which controls charge transfer between the transition metal and its adjacent atoms (C, N, O, or H), thereby regulating the binding of chemical intermediates. To describe the relationships between these two parameters (Nd and ETM), a descriptor ψ was introduced.61,62 Which is represented as,
![]() | (9) |
Higher ψ generally means weaker binding to intermediates due to more filled d-orbitals leading to stronger antibonding state occupation. Fig. 9c shows a plot of ψ against UL for TM2@BPN, with Rh2@BPN at the peak of the volcano plot which corresponds to the activity results. This suggests that ψ shows a strong linear correlation with adsorption energies and catalytic activity, providing a predictive tool for designing catalysts TM2@BPN screening of NO3RR catalysts. Using a catalyst's easily available intrinsic nature as a descriptor to forecasting its catalytic activity provides a cost-effective and practical technique, that avoids the need for costly DFT calculations and experimental experiments.
To enable a deeper understanding of reaction pathways and electron transfer mechanisms, the gradual change of charge throughout the reaction pathway of TM2@BPN for NO3RR was investigated based on Bader charge analysis. Each intermediate is subdivided into three moieties (Fig. 10a), moiety 1 corresponds to (BPN support), moiety 2 is the (TM2C4) catalytic site and moiety 3 denotes the adsorbed intermediates involved in the reaction pathway. Fig. 10b, c and S18 display the charge variation on Rh2@BPN, Os2@BPN and Ir2@BPN. Throughout the NO3RR process, all three moieties exhibit variation in charge. During the first step, the NO3− molecule receives 0.74 e−, 0.80 e−, and 0.69 e− from moiety 2 and moiety 1 on Rh2@BPN, Os2@BPN, and Ir2@BPN respectively. The moiety 2 consistently donates electrons to the intermediates, which is important for improving the adsorption of intermediates. The final increase in moiety 3's charge signifies NH3 desorption proceeds without difficulty. Thus, Rh2@BPN, Os2@BPN, and Ir2@BPN exhibit a comparable charge variation trend. As the adsorbed intermediates receive electrons, moiety 1 facilitates electron transfer between the catalytic site and the adsorbed intermediates, allowing the catalytic process to proceed continuously, moiety 2 functions as both the DAC's active site and an electron donor.
![]() | ||
| Fig. 10 (a) Three moieties of TM2@BPN. (b and c) Charge variation during the process of NO3RR on Rh2@BPN and Os2@BPN. | ||
![]() | (10) |
![]() | ||
| Fig. 12 Ab initio molecular dynamics simulation of Rh2@BPN at 500 K for a total duration of 10 ps with a time step of 2 fs. | ||
Supplementary information: Notes 1, 2 and 3 for elementary and kinetic steps of NO3RR, solvent effect, and faradaic efficiency, four possible binding configurations of TM dimer on BPN (Fig. S1), three possible adsorption configurations of NO3− on TM2@BPN (Fig. S2), DFT-optimized configurations of intermediates associated with free energy diagram of Rh2@BPN, Os2@BPN and Ir2@BPN respectively (Fig. S3), free energy diagram for NO3RR on Mo2@BPN, Ru2@BPN and Pt2@BPN catalysts respectively, (UL > −0.5 V) (Fig. S4), free energy diagram for NO3RR on Sc2@BPN, Ti2@BPN and V2@BPN catalysts respectively (Fig. S5), free energy diagram for NO3RR on Mn2@BPN, Fe2@BPN and Co2@BPN catalysts respectively (Fig. S6), free energy diagram for NO3RR on Ni2@BPN, Y2@BPN and Zr2@BPN catalysts respectively (Fig. S7), free energy diagram for NO3RR on Nb2@BPN, Pd2@BPN and Hf2@BPN catalysts respectively (Fig. S8), free energy diagram for NO3RR on Ta2@BPN, W2@BPN and Re2@BPN catalysts respectively (Fig. S9), three possible adsorption configurations of NO on TM2@BPN (Fig. S10), nudged elastic band calculation results for *NO + (H3O+ + e−) → *NOH + H2O on Rh2@BPN (Fig. S11), nudged elastic band calculation results for *NO + (H3O+ + e−) → *NOH + H2O (Fig. S12), two possible adsorption configurations of NO3− on TM@BPN (Fig. S13), free energy diagram for NO3RR on Rh@BPN, Os@BPN and Ir@BPN SACs catalysts respectively (Fig. S14), (a) charge density differences of nitrate adsorbed on Rh2@BPN, and Os2@BPN the isosurface value is 0.001 e/Å3. (b) Schematic representation of the nitrate adsorption mechanism on the TM dimer (Fig. S15), charge density differences of nitrate adsorbed on Mo2@BPN, Ru2@BPN, Ir2@BPN and Pt2@BPN, the isosurface value is 0.001 e/Å3. The charge depletion and accumulation were depicted by cyan and violet respectively (Fig. S16), partial density of states (PDOS) on the molecular orbital of NO3 and the d orbital of metal atoms on Mo2@BPN, Ru2@BPN, Ir2@BPN and Pt2@BPN catalysts after NO3 adsorption (Fig. S17), three moieties of TM2@BPN and charge variation during the process of NO3RR on Ir2@BPN (Fig. S18), ab initio molecular dynamics simulation of (a) Os2@BPN, and (b) Ir2@BPN was carried out at 500 K for 1 ps over a total time step of fs (Fig. S19). Zero-point energy (ΔEZPE) and entropy (TΔS) at room temperature of gaseous molecules from the NIST database (Table S1), the computed binding energy (Eb), cohesive energy (Ecoh) and the difference between binding and cohesive (Eb–Ecoh) of TM dimers in TM2@BPN (Table S2), electronic and structural properties of TM2@BPN, including transition metal (TM) dimer bond length (dM–M), charge transfer (CT) from TM dimer to substrate, and d band center (εd) (Table S3), calculated adsorption free energy ΔG*NO3 (eV) on TM2@BPN, via (η1-O-1), (η2-O-1) and (η2-O-2) configuration (Table S4), transition metal to oxygen bond distance for the stable NO3− adsorbed configuration (η2-O-2) on TM2@BPN (Table S5), calculated Gibbs free energy of *H (ΔG*H) adsorption on TM2@BPN (Table S6), calculated charge transfer (CT) from TM dimer to NO3−, on TM2@BPN (Table S7), the Computed limiting potential (UL) and potential determining step (PDS) of TM2@BPN for NO3RR (Table S8), the calculated Zero-point energy (ΔEZPE) and entropic contributions (TΔS) at room temperature (298.15 K) to the Gibbs free energies of each elementary step for Rh2@BPN. The unit is eV (Table S9), the calculated Zero-point energy (ΔEZPE) and entropic contributions (TΔS) at room temperature (298.15 K) to the Gibbs free energies of each elementary step for Os2@BPN. The unit is eV (Table S10), the calculated Zero-point energy (ΔEZPE) and entropic contributions (TΔS) at room temperature (298.15 K) to the Gibbs free energies of each elementary step for Ir2@BPN. The unit is eV (Table S11),calculated adsorption free energy ΔG*NO (eV) on TM2@BPN, via *NO end, *ON end, and **NO side configurations (Table S12), calculated adsorption free energy ΔG*NO3 (eV) on TM@BPN, via 1-O, and 2-O configuration (Table S13), the comparison of catalytic performance (reflected by limiting potential, UL) for Nitrate reduction reaction on various electrocatalysts (Table S14), calculated d-band center (εd) of TM@BPN (Table S15), calculated limiting potential of HER and NO3RR and faradaic efficiency (FE) values for TM2@BPN (Table S16), comparison of nitrate reduction reaction pathways on the Rh2@BPN by using DFT-D3 and DFT-Sol method (Table S17), comparison of nitrate reduction reaction pathways on the Os2@BPN by using DFT-D3 and DFT-Sol method (Table S18). See DOI: https://doi.org/10.1039/d5ta07089a.
| This journal is © The Royal Society of Chemistry 2026 |