Next Article in Journal
Mathematical Modeling of the Kinetics of Glucose Production by Batch Enzymatic Hydrolysis from Algal Biomass
Previous Article in Journal
Photothermal Effect of Carbon-Doped Carbon Nitride Synergized with Localized Surface Plasmon Resonance of Ag Nanoparticles for Efficient CO2 Photoreduction
Previous Article in Special Issue
Controlled Synthesis of Nickel Phosphides in Hollow N, P Co-Doped Carbon: In Situ Transition to (Oxy)hydroxide Phases During Oxygen Evolution Reaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synergistic Effect of the Heteronuclear Double Sites in C9N4 on the Electrochemical Reduction of CO2 to CO

College of Science, University of Shanghai for Science and Technology, Shanghai 200093, China
*
Author to whom correspondence should be addressed.
Catalysts 2025, 15(4), 370; https://doi.org/10.3390/catal15040370
Submission received: 20 March 2025 / Revised: 6 April 2025 / Accepted: 8 April 2025 / Published: 10 April 2025
(This article belongs to the Special Issue Recent Advances in Electrocatalysis and Future Perspective)

Abstract

:
In response to the detrimental impact of excessive fossil fuel usage on the environment and the looming energy crisis, the electrochemical reduction of carbon dioxide (CO2RR) has emerged as a promising solution. This study investigates the potential of dual-atom catalysts, specifically boron (B) and transition metal (TM) co-modified C9N4, for efficient CO2RR. The 2 × 2 × 1 supercell of C9N4, considering modification with 26 TM and B atoms, demonstrated stability, confirmed by binding and formation energy calculations. Molecular dynamics simulations further supported the thermal stability of the studied catalysts. The modified structures exhibited metallic behavior, suggesting potential facilitation of electron transfer during electroreduction. Furthermore, by conducting Gibbs free energy calculations on CO2 reduction pathways, seven low overpotential catalysts were screened out. Considering the competitive hydrogen evolution reaction (HER), Sc-B and Hf-B demonstrate excellent selectivity towards CO2, with Faradaic efficiencies (FE) close to 100%, and possess low limiting potentials of −0.30 and −0.53 eV, showcasing their potential to be excellent catalysts. The introduction of pre-adsorbed hydrogen atoms further optimized the advantage of CO2RR over HER, with the efficiencies of Ti-B@C9N4-H and Hf-B@C9N4-H methods increasing from 0% and 28% to over 99%, respectively, providing new insights into overcoming the low selectivity of CO2 reduction.

1. Introduction

In recent years, with the rapid development of the global economy, the excessive use of fossil fuels has had adverse effects on people’s lives and development. Large-scale emissions of carbon dioxide exacerbate the greenhouse effect. Secondly, an energy crisis is gradually looming [1,2]. To address these challenges, the use of clean energy sources such as solar and wind power to convert into electricity and conduct CO2 reduction reactions (CO2RR) under mild conditions has become an increasingly mature method [3,4,5,6,7]. This approach involves the electrochemical reduction of excess CO2 from the atmosphere to generate industrially valuable C1, C2, or high-carbon products, such as carbon monoxide, formic acid, methane, ethanol, acetic acid, etc. [8,9,10,11,12,13,14]. CO exhibits application value in chemical synthesis (as a reducing agent), materials science (as a refrigerant), and the fuel sector [15,16]. So far, suitable catalysts for electrocatalytic CO2 reduction to CO are being explored, conducting extensive studies in both experimental and theoretical aspects. Ren et al. proposed that the Ni/Fe-NC dual-atom site catalyst (DAC) exhibited good CO2RR performance in CO production and that the CO selectivity could remain above 90% over a wide potential range from −0.5 to −0.9 V [17]. Beyond that, Zhang et al. experimentally demonstrated that a supported Pd2 DAC was used for electrochemical CO2RR. It exhibited superior CO2RR catalytic performance with 98.2 % CO Faradic efficiency at −0.85 V vs. RHE [18]. As a product, it demonstrates a relatively high Faradaic efficiency.
With the successful isolation of monolayer graphene [19,20], the investigated graphene-based nano-meshes (GNMs) with nanopore arrays show superior electronic properties to graphene. For example, graphitic carbon nitride (g-C3N4) has been synthesized experimentally for decades and is still being investigated today. Zhu et al. investigated g-C3N4-based single-atom catalysts (SACs), Ti-, Ag-, and Cr-gC3N4 with low limiting potentials for producing different C1 products. In addition, Liu et al. reported that the nitrogen-coordinated graphene substrate composed of Cu and different TM double-doped atoms enhances the catalytic performance for producing C1 products [21].
The theoretically predicted synthesis of C9N4 has shown a two-dimensional honeycomb-kagome lattice near the Fermi level [22,23]. Its porous environment and nitrogen coordination facilitate the anchoring of TMs and prevent aggregation, making it widely applicable in catalysts and topological materials. SACs based on C9N4 have been studied [24]; however, there is still room for improvement in the overpotential and Faradaic efficiency in the electrochemical CO2RR on SAC systems. Introducing a new type of metal atom into SACs to form bimetallic catalytic sites is an effective method to enhance the catalytic performance of CO2RR. However, achieving synergistic interactions between different atoms within the same carrier remains a challenge [25,26,27]. In previous reports, B atoms have been proposed as active centers for CO2 electrocatalytic reduction [28]. The sp3 orbitals of B share similar functionalities with the d orbitals of TMs, and B@g-C9N4 materials can effectively suppress the HER reaction [29]. Therefore, the use of B and TM atoms as dual-atom catalysts is a promising strategy. By introducing dual-active sites, better coordination of interactions between atoms can be achieved, regulating the adsorption of intermediates and controlling catalytic activity and selectivity [30,31,32,33]. This strategy holds promise for providing a new pathway for electrochemical CO2 reduction, laying the foundation for achieving high catalytic performance.
In this study, C9N4 was chosen as the substrate, and B, along with 26 TM co-modified catalysts, was investigated. A series of B-TM catalysts were studied for their stability, catalytic activity, and selectivity in CO2RR. Two catalysts (Sc-B@C9N4, Hf-B@C9N4) with high reaction activity and selectivity in reducing CO2 to CO were identified after screening. Subsequently, modifications were made to systems prone to competitive hydrogen evolution reactions, yielding promising results that could provide new insights for improving the low selectivity of CO2 reduction.

2. Results and Discussion

2.1. The Structure and Stability of TM-B@C9N4

The selection of a 2 × 2 × 1 supercell of C9N4 was initiated to provide adequate voids for accommodating TM and B atoms while minimizing the impact of other atoms on the calculations (Figure 1a).
The stability of the substrate material is a crucial factor to consider for excellent catalysts, as it directly influences the catalyst’s usage lifespan. To address this concern, the binding energies (Eb) and formation energies (Ef) of 26 TM atoms anchored in B@C9N4 pores were calculated using the formula (Figure 1b). Eb < 0 indicates stable binding of TM atoms to the substrate, while Ef < 0 suggests the feasibility of experimental synthesis. Except for Os atoms not meeting the criteria (Ef > 0), highly negative binding energies and relatively negative formation energies confirm the stability of the structure. The optimized substrate structure is shown in Figure S1. Molecular dynamics simulations (Figure S2), using Sc-B@C9N4 as an example, were employed to further evaluate the thermal stability of the studied catalysts. At a temperature of 300 K and a time step of 3 fs, no significant deformation or dissociation of the geometric structure was observed within 12 ps. Temperature and energy exhibited minor fluctuations within a small range, indicating the strong stability of this catalyst, warranting further investigation.
Based on the selection of substrate catalytic activity described in the article, TM-B@C9N4 (TM = Sc\Ti\V\Mn\Zr\Ag\Hf) was chosen as the object of analysis. We calculated the electron localization function (ELF) to investigate the bonding characteristics of the substrate (Figure 1c) [34,35,36]. The gradient-colored image represents different degrees of electron localization: red (value 1) signifies complete electron localization, blue (value 0) indicates complete electron delocalization, and green (value 0.5) corresponds to a state resembling a free electron gas. Significant electron localization was observed near the N atoms within the cavity, which aids in anchoring metal atoms. The bonding between B atoms and coordinating N atoms exhibited covalent characteristics, highlighting a strong electron-sharing interaction. The ELF of the other six substrates also exhibited similar electronic characteristics and bonding patterns (as shown in Figure S3), demonstrating the rationality and stability of the TM-B@C9N4 structure. Additionally, the ELF diagram indicates charge transfer between the TM and B atoms, which is further confirmed by the partial density of states (PDOS) graph in Figure S4. The graph shows coupling between the d orbitals of the TM atoms in the substrate and the p orbitals of the B atoms, suggesting an interaction between TM and B atoms that facilitates charge transfer and redistribution, which is beneficial for subsequent CO2 adsorption and activation.
Additionally, the band structure of the TM-B@C9N4 substrate was studied (Figure 1d) and compared to the original C9N4 [23]; the conduction band bottom crosses the Fermi level, exhibiting metallic properties. This electronic characteristic is expected to facilitate electron transfer during the electroreduction process. The band structures of the remaining tested structures also exhibit similar results (Figure S5). In conclusion, C9N4 modified with TM-B atoms demonstrates favorable stability and unique electronic properties, holding the potential to become an excellent catalyst.

2.2. Adsorption and Activation of CO2

The initial step in previous CO2RR processes involves the adsorption and activation of CO2, forming the foundation for subsequent CO2 hydrogenation. Consequently, the structural optimization of CO2 adsorption on different substrates was conducted, yielding various adsorption configurations (Figure S6). Taking Sc-B@C9N4 as an example (Figure 2a), C and O atoms form bonds with TM and B atoms, causing the linear CO2 molecule to undergo a V-shaped distortion with an ∠OCO angle of 127.70°. The C-O bond lengths are elongated by 0.147 Å and 0.069 Å compared to free CO2 gas molecules, indicating CO2 bond activation. Similar V-shaped distortions were observed for CO2 adsorbed on other substrates (Table S1), confirming CO2 bond activation.
Furthermore, differential charge density and Bader charge calculations were employed to better analyze the electronic characteristics of CO2 adsorption and how electrons are transferred from dual-active centers to activate CO2. As shown in Figure 2a and Table S2, after CO2 adsorption, Sc and B atoms lose 0.02 and 0.64, respectively. After CO2 adsorption, Sc loses fewer electrons because, as a transition metal with low electronegativity and vacant d orbitals, it acts as a Lewis acid, attracting the lone pair electrons from oxygen and bonding with the oxygen atom to form a metal–oxygen bond. In contrast, B atoms donate more electrons due to their slightly lower electronegativity compared to C. As an electron acceptor, C interacts with the electrons donated by B, forming a stable B-C covalent bond. Therefore, in the interaction with CO2, the TM and B atoms function as a Lewis acid-base pair, promoting electron transfer and molecular activation. The remaining C9N4 substrate acts as an electron reservoir, transferring electrons from the active centers to the CO2 molecule, and it loses 0.27. Hence, CO2 gains a total of 0.94, facilitating subsequent hydrogenation reactions. Similar changes are observed in the adsorption of CO2 in the other six systems depicted in Figure S7.
We also examined the projected density of states (PDOSs) of the CO2 adsorption intermediates (Figure 2b). From the PDOSs obtained after CO2 adsorption on the substrate, it is evident that there exists a strong coupling between the d orbitals of Sc and the p orbitals of B with the p orbitals of CO2. The simultaneous downward shift of bonding orbitals below the Fermi level indicates the cooperative stabilization of CO2 by Sc and B atoms. Simultaneously, the partially occupied orbitals of Sc and B provide electrons to the 2p * orbitals of CO2, causing some partially occupied 2p * orbitals to cross the Fermi level, indicating effective activation of C-O bonds for the subsequent hydrogenation. Similar changes were observed in other systems with CO2 adsorption (Figure S8).
Finally, the projection of the crystal orbital Hamiltonian population (PCOHP) for the C-O bond in CO2, both before and after adsorption, was calculated (Figure 2c). Of the two C-O bonds in CO2, focus was placed on the C-O bond relevant to the subsequent hydrogenation step. The integration of bonding and antibonding states below the Fermi level yields the integral crystal orbital Hamiltonian population (ICOHP) values, indicating the strength of chemical bond formation. More negative ICOHP values indicate stronger bonding. The ICOHP value of the CO bond in the free CO2 gas molecule is approximately −18.45 eV. After adsorption on the Sc-B@C9N4 substrate, the value increases to −15.80 eV, indicating that the CO bond is effectively weakened and activated. In Figures S8 and S9, the adsorption of CO2 onto other TM-B@C9N4 (TM = Sc\Ti\V\Mn\Zr\Ag\Hf) substrates induces positive changes in ICOHP values, specifically, −15.94, −15.46, −15.11, −15.06, −13.4, and −16.27 eV, respectively, further confirming that TM and B-modified C9N4 effectively activate C-O bonds, laying the foundation for subsequent CO2 reduction.

2.3. The Mechanism and Free Energy of CO2 Reduction to CO

The proposed CO2RR pathway is illustrated in Figure 3a. It is observed that CO2 molecules, after adsorption, can undergo attack by a proton/electron pair to form *COOH species. Subsequently, the *COOH species gains a proton/electron pair to form *CO species, releasing a water molecule. Finally, the *CO species desorbs from the substrate. Following this pathway, Gibbs free energy calculations were performed for the selected catalysts, and the maximum free energy step differences during the reduction process were recorded (Figure 3b). A catalyst was considered potentially efficient if the absolute value of the limiting potential was less than 0.61 [37,38]. Accordingly, seven catalysts, Sc\Ti\V\Mn\Zr\Ag\Hf-B@C9N4, were screened. Additionally, we observed that systems from different periods exhibit similar trends, possibly related to the outermost electron count of different TM atoms.
To investigate the potential determining steps for CO2 reduction to CO and differences in activity on various catalysts, Gibbs free energy diagrams were plotted for the seven mentioned catalysts, as shown in Figure 4. Among these, on Ti-B@C9N4, the energy required for the first step of CO2 hydrogenation to form *COOH species is 0 eV, while the uphill energy needed for the second pair (H+/e-) attacking *COOH and dehydrating is 0.26 eV. Finally, the downhill energy for CO desorption is −0.05 eV. Therefore, the potential-determining step is the second hydrogenation step to produce *CO, with a corresponding limiting potential of −0.26 eV. Except for Mn-B@C9N4 and V-B@C9N4, where the potential-determining step is the first hydrogenation step producing *COOH, the remaining four catalysts involve the second hydrogenation step. The desorption energies of CO for Ti, V, Mn, and Ag-B@C9N4 catalysts are all lower, indicating spontaneous desorption. Sc, Zr, and Hf-B@C9N4 require only a small increase in energy, suggesting that they have potential as high-efficiency catalysts. Due to the relatively low free energy step of *+CO (0.21 V), the difference between the free energy steps of CO and +CO reflects the ease of CO desorption. For example, Wu et al. employed a graphene-supported nickel hydride cluster (7HNi10-gra) as a catalyst to precisely tune the external potential, thereby regulating the concentrations of formate and methanol intermediates. This approach allowed them to rationally predict the formation of formic acid, methanol, and methane as the three main products. In the CO2RR process reported in their work, the Gibbs free energy step for *CO is relatively low, which favors further hydrogenation to form other hydrocarbons. In contrast, on TM-B@C9N4 (TM = Sc\Ti\V\Mn\Zr\Ag\Hf), the *CO free energy step is significantly higher, promoting CO (g) desorption as the dominant pathway [39].

2.4. Competitive HER and Catalyst Improvements

Generally, catalysts undergo CO2 reduction in aqueous solutions. To determine the efficiency of obtaining the desired products, it is necessary to consider the competition with hydrogen evolution reaction (HER), as both CO2RR and HER significantly consume proton/electron pairs and share active sites. The occurrence of HER can lead to a decrease in the Faradaic efficiency of CO2RR. Therefore, the Gibbs free energy of HER on TM-B@C9N4 (TM = Sc, Ti, V, Mn, Zr, Ag, Hf) was initially calculated (Figure 5a), and the corresponding limiting potentials are provided in Table S3. Additionally, the adsorption energies of CO2 and H were considered, filtering out catalysts with excellent selectivity. As depicted in Figure 5b, this implies that the catalyst predominantly favors the reduction of CO2 during hydrogenation [40]. The more positive this value, the higher the selectivity of CO2 reduction over HER. Sc, Ag, and Hf-B@C9N4 all have values greater than zero. The competitively unfavorable process during CO2RR is the hydrogen evolution reaction (HER), which has a significant negative influence on CO2RR selectivity. As a result, the following assumptions are made to evaluate the selectivity of CO2RR:
(1)
HER is the only side reaction in the CO2RR process;
(2)
The rate of HER and CO2RR do not depend mainly on proton and electron transfer;
(3)
Boltzmann distribution can be used to estimate the selectivity of CO2RR relative to HER.
The theoretical Faradaic efficiency (FE) of CO2RR can be calculated according to the Boltzmann distribution:
F E = 1 1 + e Δ G k B T × 100 %
where ∆G is the Gibbs free energy difference between the HER and CO2RR potential determination step, kB represents the Boltzmann constant, and T = 298.15 K [41]. The efficiencies of the three catalysts are all above 99%. While Zr-B@C9N4 does not exhibit a clear preference for CO2RR on CO2 desorption, it still shows a 28% Faradaic efficiency (Table S3). The more negative CO2 adsorption energy compared to H also indicates the substrate’s preference for CO2RR. However, it is not satisfied on the substrate Ag-B@C9N4. Therefore, only Sc and Hf-B@C9N4 meet the requirements to suppress HER.
Moreover, during practical electrocatalytic processes, significant discrepancies may arise between the local pH near the catalyst surface and the bulk electrolyte environment due to proton consumption and limited diffusion. This local pH gradient can influence proton-coupled electron transfer (PCET) steps, thereby altering the thermodynamic and kinetic competition between the CO2RR and the HER. Previous studies have demonstrated that under alkaline or high-pH conditions, the formation of *H intermediates is thermodynamically less favorable, which helps suppress HER. Simultaneously, key intermediates along the CO pathway, such as *COOH, become more stabilized, further enhancing the selectivity toward CO production [6]. Although the standard computational hydrogen electrode (CHE) model at pH = 0 was adopted in this study without explicitly considering local pH effects, the thermodynamic data and electronic structure analysis still indicate that the proposed heteronuclear dual sites exhibit favorable CO selectivity. Future investigations could incorporate explicit solvation models or potential-dependent simulations to gain deeper insights into the regulatory role of the local environment on catalytic mechanisms.
Figure 5. (a) Hydrogen evolution reaction (HER) free energy diagrams on these seven catalysts. (b) The difference in limit potential (UL) between CO2 reduction to CO and hydrogen evolution reaction to H2 on TM-B@C9N4, as well as the changes in CO2 adsorption energy and H adsorption energy (* represents substrate material).
Figure 5. (a) Hydrogen evolution reaction (HER) free energy diagrams on these seven catalysts. (b) The difference in limit potential (UL) between CO2 reduction to CO and hydrogen evolution reaction to H2 on TM-B@C9N4, as well as the changes in CO2 adsorption energy and H adsorption energy (* represents substrate material).
Catalysts 15 00370 g005
There are localized electrons around the N atoms in the substrate cavity, which may adsorb H atoms. Ti, V, and Zr-B@C9N4 also have the potential to act as highly selective catalysts. Therefore, the effect on selectivity and activity was explored by pre-adsorbing H atoms onto N atoms not bonded to TM and B in the above materials (Figure 6a). The adsorption energy of CO2, the limiting potential for CO2RR, and the Gibbs free energy for HER on Ti, V, and Zr-B@C9N4-H were calculated (as shown in Table S4). It was found that, after pre-adsorbing H, the CO2 adsorption energy remained almost unchanged, while the limiting potential for CO2RR showed varying degrees of change. Except for V-B@C9N4, which exhibited a small change, the absolute values of the limiting potentials for Ti-B@C9N4-H and Zr-B@C9N4-H decreased to 0.19 eV and 0.41 eV, respectively. The corresponding Gibbs free energy diagram is shown in S10. Furthermore, the pre-adsorption of H on Ti/V-B@C9N4 led to a lower limiting potential for HER, hindering its progression. The corresponding H adsorption energy decreased; however, it remained higher than the CO2 adsorption energy. HER on Zr-B@C9N4-H was almost unaffected. It is noteworthy that, during the structural optimization process, there was no displacement observed in the H atom, indicating its stability. Through Figure 5b, it can be visually confirmed that both Ti and Zr-B@C9N4-H satisfy and indicate the predominance and priority of CO2RR.
Bader charge calculations for the substrate with added H to the cavity were performed and compared with the base substrate without added H (Table 1). It was found that in the system with pre-adsorbed H, both TM and B active centers lose fewer electrons when binding to the substrate because H atoms share some electrons with the substrate. This suggests that more electrons may be involved in CO2 adsorption and activation in subsequent steps, lowering the barrier for CO2RR. This approach could provide a new perspective for improving the low selectivity of CO2.

3. Computational Methodology

All computations were carried out utilizing the spin-polarized density functional theory (DFT) methodology implemented in the Vienna ab initio simulation package (VASP 5.4.4) [42]. The electron–ion interactions were modeled using the projector-augmented wave (PAW) approach, while the electronic exchange and correlation effects were described using the Perdew–Burke–Ernzerhof (PBE) functional within the framework of the generalized gradient approximation (GGA) [43,44]. To account for the van der Waals (vdW) interactions between reactants, intermediates, and catalysts, the empirical correction scheme based on the Grimme method (DFT+D3) was employed [45]. A plane wave energy cutoff of 400 eV was employed for structural optimizations, while a value of 500 eV was chosen for electronic structure calculations. The k-point grids were set to 1 × 1 × 1 for geometry optimization and 6 × 6 × 1 for electronic structure calculations. Convergence criteria for electronic energy and forces were established at 10–5 eV and 0.05 eV/Å, respectively. Additionally, an 18 Å vacuum layer was introduced to mitigate interlayer interactions. Catalyst thermal stability was assessed through ab initio molecular dynamics (AIMD) simulations using VASP, with a Nosé–Hoover thermostat employed. A time step of 3 fs and a total simulation duration of 12 ps were selected to verify catalyst thermal stability. To ascertain the Gibbs free energy during proton transfer in CO2RR, computations were conducted utilizing a computational hydrogen electrode (CHE) model.
Gibbs free energy changes (ΔG) for all intermediate steps were determined using the following equation: Δ G = Δ E + Δ Z P E T Δ S + Δ G p H , where ΔE corresponds to the total energy of surface and intermediate species obtained from DFT calculations, ΔZPE represents the variation in zero-point energy, and ΔS accounts for the entropy change of free molecules and adsorbates at 298.15 K. The pH-dependent ΔG values were calculated using the formula: Δ G p H = k B T × l n 10 × p H , where kB is the Boltzmann constant, and pH was set to zero.
The calculation of UL is determined by the following formula:   U L = Δ G m a x / e , here, ΔGmax is the maximum Gibbs free energy change in the reaction. The negative sign indicates the reduction potential, and a smaller |UL| signifies higher catalytic activity.
In order to evaluate the stability of the new structure after doping of transition metal (TM), the binding energy (Eb) of the doped metal atoms was calculated using the following equation: E b = E TM B @ C 9 N 4 E C 9 N 4 E TM single E B where E TM B @ C 9 N 4 represents the total energy associated with the TM and B atom modifications on C9N4, denotes the energy of the pristine C9N4 substrate structure, and ETM-single and EB, respectively, denote the energies associated with the transition metal (TM) atom and the non-metallic boron (B) atom.
Adsorption energy refers to the energy released or consumed when CO2 adsorbs onto the catalyst surface. It is calculated using the following formula:   E a d s = E t o t a l ( E c a t a l y s t + E a d s o r b a t e ), where Eads is the total energy of the adsorbate on the catalyst surface, where Ecatalyst is the total energy of the catalyst itself, and Eadsorbate is the total energy of the adsorbate.

4. Conclusions

This study employed C9N4 as the substrate, utilizing 26 transition metal (TM) atoms along with boron (B) atoms as dual active sites for CO2 reduction reactions (CO2RR). The calculations of binding energy (Eb), formation energy (Ef), and ab initio molecular dynamics (AIMD) revealed that the majority of TM-B@C9N4 systems exhibit excellent thermal stability. The heteronuclear dual active sites synergistically enhance CO2 adsorption and activation, as elucidated through an in-depth analysis of charge density differences, electron localization function (ELF), local density of states (LDOS), and integrated crystal orbital Hamilton population (ICOHP) for selected systems. Gibbs free energy calculations identified six catalysts, namely, Sc-B@C9N4, Hf-B@C9N4, Ti-B@C9N4, V-B@C9N4, Mn-B@C9N4, and Zr-B@C9N4 Ag-B@C9N4, exhibiting low overpotentials for CO generation (UL < 0.61). Considering the suppression of competitive hydrogen evolution reactions (HERs), only Sc-B@C9N4 and Hf-B@C9N4 displayed superior selectivity, achieving nearly 100% Faradaic efficiency. Modification of less selective materials resulted in improved selectivity for Ti-B@C9N4-H and Zr-B@C9N4-H upon hydrogenation, accompanied by minor changes in overpotential. Insight into this phenomenon was gained through Bader charge analysis. This theoretical investigation provides new strategies for the design and synthesis of high-performance catalysts while offering novel insights into the mechanism of CO2 activation. It is anticipated that this research will serve as a valuable reference for the further development of novel CO2RR electrocatalysts.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal15040370/s1, Figure S1: 26 optimized TM-B@C9N4 structures; Figure S2: The variation of energy and temperature with respect to the simulation time for Sc-B@C9N4 at 300 K in AIMD, with a time step of 12 ps. The figures display the top and side views of Sc-B@C9N4 at the final AIMD step; Figure S3: Electronic localization function image of Ti-B@C9N4, V-B@C9N4, Hf-B@C9N4, Mn-B@C9N4, Zr-B@C9N4, and Ag-B@C9N4; Figure S4: PDOS (Partial Density of States) plots for the d orbitals of TM atoms and the p orbitals of B atoms on seven TM-B@C9N4 substrates; Figure S5: Band map of six TM-B@C9N4; Figure S6: Pmized adsorption configurations of CO2 on the 26 TM-B@C9N4; Figure S7: Difference charge density of CO2 molecule adsorbed on six TM-B@C9N4. Charge depletion and accumulation are presented in cyan and yellow, respectively. The value of isosurface is 0.008 e/Å3; Figure S8: Comparison of the p-orbitals of CO2 and B and the d-orbitals of TM after CO2 adsorption on the six TM-B@C9N4 systems; Figure S9: PCOHP and bond length between C and O atoms of CO2 adsorbed on six TM-B@C9N4 systems; Figure S10: Comparison of the free energy distribution for three TM-B@C9N4 and TM-B@C9N4-H configurations under zero applied voltage (vs. RHE); Table S1: C-O bond lengths and bond angles of CO2 adsorbed on six TM-B@C9N4 substrates; Table S2: Charge transfer amounts for six TM-B@C9N4 substrates after CO2 adsorption, where negative numbers represent electron loss and positive signs represent electron acquisition; Table S3: The limiting potentials for six types of TM-B@C9N4 systems in the CO2RR and HER processes, adsorption energy information for CO2 and H, as well as the Faradaic efficiency of CO2RR relative to HER; Table S4: The limiting potentials for three types of TM-B@C9N4-H systems in the CO2RR and HER processes, adsorption energy information for CO2 and H, as well as the Faradaic efficiency of CO2RR relative to HER.

Author Contributions

Writing—original draft, R.W. and B.Z.; Writing—review & editing, Z.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data supporting this article have been included as part of the Supplementary Information.

Acknowledgments

This work is supported by the High Performance Computing Center, University of Shanghai for Science and Technology.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Rudd, J.A.; Hernandez-Aldave, S.; Kazimierska, E.; Hamdy, L.B.; Bain, O.J.E.; Barron, A.R.; Andreoli, E. Investigation into the Re-Arrangement of Copper Foams Pre-and Post-CO2 Electrocatalysis. Chemistry 2021, 3, 687–703. [Google Scholar] [CrossRef]
  2. Chu, S.; Cui, Y.; Liu, N. The Path towards Sustainable Energy. Nat. Mater. 2017, 16, 16–22. [Google Scholar] [CrossRef]
  3. Costentin, C.; Robert, M.; Savéant, J.-M. Catalysis of the Electrochemical Reduction of Carbon Dioxide. Chem. Soc. Rev. 2013, 42, 2423–2436. [Google Scholar] [CrossRef] [PubMed]
  4. Kuhl, K.P.; Cave, E.R.; Abram, D.N.; Jaramillo, T.F. New Insights into the Electrochemical Reduction of Carbon Dioxide on Metallic Copper Surfaces. Energy Environ. Sci. 2012, 5, 7050–7059. [Google Scholar] [CrossRef]
  5. Zhu, D.D.; Liu, J.L.; Qiao, S.Z. Recent Advances in Inorganic Heterogeneous Electrocatalysts for Reduction of Carbon Dioxide. Adv. Mater. 2016, 28, 3423–3452. [Google Scholar] [CrossRef]
  6. Gao, D.; Arán-Ais, R.M.; Jeon, H.S.; Roldan Cuenya, B. Rational Catalyst and Electrolyte Design for CO2 Electroreduction towards Multicarbon Products. Nat. Catal. 2019, 2, 198–210. [Google Scholar] [CrossRef]
  7. Ross, M.B.; De Luna, P.; Li, Y.; Dinh, C.-T.; Kim, D.; Yang, P.; Sargent, E.H. Designing Materials for Electrochemical Carbon Dioxide Recycling. Nat. Catal. 2019, 2, 648–658. [Google Scholar] [CrossRef]
  8. Kornienko, N.; Zhao, Y.; Kley, C.S.; Zhu, C.; Kim, D.; Lin, S.; Chang, C.J.; Yaghi, O.M.; Yang, P. Metal-Organic Frameworks for Electrocatalytic Reduction of Carbon Dioxide. J. Am. Chem. Soc. 2015, 137, 14129–14135. [Google Scholar] [CrossRef]
  9. Ma, W.; Xie, S.; Zhang, X.-G.; Sun, F.; Kang, J.; Jiang, Z.; Zhang, Q.; Wu, D.-Y.; Wang, Y. Promoting Electrocatalytic CO2 Reduction to Formate via Sulfur-Boosting Water Activation on Indium Surfaces. Nat. Commun. 2019, 10, 892. [Google Scholar] [CrossRef]
  10. Shen, J.; Kortlever, R.; Kas, R.; Birdja, Y.Y.; Diaz-Morales, O.; Kwon, Y.; Ledezma-Yanez, I.; Schouten, K.J.P.; Mul, G.; Koper, M.T. Electrocatalytic Reduction of Carbon Dioxide to Carbon Monoxide and Methane at an Immobilized Cobalt Protoporphyrin. Nat. Commun. 2015, 6, 8177. [Google Scholar] [CrossRef]
  11. Zhang, Y.; Liu, J.; Wei, Z.; Liu, Q.; Wang, C.; Ma, J. Electrochemical CO2 Reduction over Nitrogen-Doped SnO2 Crystal Surfaces. J. Energy Chem. 2019, 33, 22–30. [Google Scholar] [CrossRef]
  12. Zhang, Y.; Mo, Y.; Cao, Z. Rational Design of Main Group Metal-Embedded Nitrogen-Doped Carbon Materials as Frustrated Lewis Pair Catalysts for CO2 Hydrogenation to Formic Acid. ACS Appl. Mater. Interfaces 2021, 14, 1002–1014. [Google Scholar] [CrossRef]
  13. Li, Y.; Su, H.; Chan, S.H.; Sun, Q. CO2 Electroreduction Performance of Transition Metal Dimers Supported on Graphene: A Theoretical Study. ACS Catal. 2015, 5, 6658–6664. [Google Scholar] [CrossRef]
  14. Lin, L.; Liu, T.; Xiao, J.; Li, H.; Wei, P.; Gao, D.; Nan, B.; Si, R.; Wang, G.; Bao, X. Enhancing CO2 Electroreduction to Methane with a Cobalt Phthalocyanine and Zinc-Nitrogen-Carbon Tandem Catalyst. Angew. Chem. Int. Ed. 2020, 59, 22408–22413. [Google Scholar] [CrossRef] [PubMed]
  15. Ma, S.; Kenis, P.J.A. Electrochemical Conversion of CO2 to Useful Chemicals: Current Status, Remaining Challenges, and Future Opportunities. Curr. Opin. Chem. Eng. 2013, 2, 191–199. [Google Scholar]
  16. Ganesh, I. Applications of Nanomaterials; Elsevier: Amsterdam, The Netherlands, 2018; pp. 83–131. [Google Scholar]
  17. Ren, W.; Tan, X.; Yang, W.; Jia, C.; Xu, S.; Wang, K.; Smith, S.C.; Zhao, C. Isolated Diatomic Ni-Fe Metal-Nitrogen Sites for Synergistic Electroreduction of CO2. Angew. Chem. Int. Ed. 2019, 58, 6972–6976. [Google Scholar] [CrossRef]
  18. Zhang, N.; Zhang, X.; Kang, Y.; Ye, C.; Jin, R.; Yan, H.; Lin, R.; Yang, J.; Xu, Q.; Wang, Y. A Supported Pd2 Dual-Atom Site Catalyst for Efficient Electrochemical CO2 Reduction. Angew. Chem. 2021, 133, 13500–13505. [Google Scholar] [CrossRef]
  19. Zhang, C.; He, Q.; Chu, W.; Zhao, Y. Transition Metals Doped Borophene-Graphene Heterostructure for Robust Polysulfide Anchoring: A First Principle Study. Appl. Surf. Sci. 2020, 534, 147575. [Google Scholar] [CrossRef]
  20. Duplock, E.J.; Scheffler, M.; Lindan, P.J.D. Hallmark of Perfect Graphene. Phys. Rev. Lett. 2004, 92, 225502. [Google Scholar] [CrossRef]
  21. Zhao, Q.; Crespo-Otero, R.; Di Tommaso, D. The Role of Copper in Enhancing the Performance of Heteronuclear Diatomic Catalysts for the Electrochemical CO2 Conversion to C1 Chemicals. J. Energy Chem. 2023, 85, 490–500. [Google Scholar] [CrossRef]
  22. Mortazavi, B.; Shahrokhi, M.; Shapeev, A.V.; Rabczuk, T.; Zhuang, X. Prediction of C7N6 and C9N4: Stable and Strong Porous Carbon-Nitride Nanosheets with Attractive Electronic and Optical Properties. J. Mater. Chem. C 2019, 7, 10908–10917. [Google Scholar] [CrossRef]
  23. Lu, Y.; Fan, X.; Ma, X.; Liu, J.; Li, Y.; Zhao, M. Tunable Topological Electronic States in the Honeycomb-Kagome Lattices of Nitrogen/Oxygen-Doped Graphene Nanomeshes. Nanoscale Adv. 2022, 4, 2201–2207. [Google Scholar] [CrossRef] [PubMed]
  24. Chen, J.-L.; Hu, H.-J.; Wei, S.-H. Transition Metal Anchored on C9N4 as a Single-Atom Catalyst for CO2 Hydrogenation: A First-Principles Study. Chin. Phys. B 2022, 31, 107306. [Google Scholar] [CrossRef]
  25. Fu, J.; Dong, J.; Si, R.; Sun, K.; Zhang, J.; Li, M.; Yu, N.; Zhang, B.; Humphrey, M.G.; Fu, Q.; et al. Synergistic Effects for Enhanced Catalysis in a Dual Single-Atom Catalyst. ACS Catal. 2021, 11, 1952–1961. [Google Scholar] [CrossRef]
  26. Zhang, L.; Si, R.; Liu, H.; Chen, N.; Wang, Q.; Adair, K.; Wang, Z.; Chen, J.; Song, Z.; Li, J. Atomic Layer Deposited Pt-Ru Dual-Metal Dimers and Identifying Their Active Sites for Hydrogen Evolution Reaction. Nat. Commun. 2019, 10, 4936. [Google Scholar] [CrossRef]
  27. Zheng, T.; Liu, C.; Guo, C.; Zhang, M.; Li, X.; Jiang, Q.; Xue, W.; Li, H.; Li, A.; Pao, C.-W.; et al. Copper-Catalysed Exclusive CO2 to Pure Formic Acid Conversion via Single-Atom Alloying. Nat. Nanotechnol. 2021, 16, 1386–1393. [Google Scholar] [CrossRef]
  28. Légaré, M.-A.; Bélanger-Chabot, G.; Dewhurst, R.D.; Welz, E.; Krummenacher, I.; Engels, B.; Braunschweig, H. Nitrogen Fixation and Reduction at Boron. Science 2018, 359, 896–900. [Google Scholar] [CrossRef]
  29. Yu, X.; Han, P.; Wei, Z.; Huang, L.; Gu, Z.; Peng, S.; Ma, J.; Zheng, G. Boron-Doped Graphene for Electrocatalytic N2 Reduction. Joule 2018, 2, 1610–1622. [Google Scholar] [CrossRef]
  30. An, B.; Li, Z.; Song, Y.; Zhang, J.; Zeng, L.; Wang, C.; Lin, W. Cooperative Copper Centres in a Metal-Organic Framework for Selective Conversion of CO2 to Ethanol. Nat. Catal. 2019, 2, 709–717. [Google Scholar] [CrossRef]
  31. Jiao, J.; Lin, R.; Liu, S.; Cheong, W.-C.; Zhang, C.; Chen, Z.; Pan, Y.; Tang, J.; Wu, K.; Hung, S.-F.; et al. Copper Atom-Pair Catalyst Anchored on Alloy Nanowires for Selective and Efficient Electrochemical Reduction of CO2. Nat. Chem. 2019, 11, 222–228. [Google Scholar] [CrossRef]
  32. Huang, Q.; Liu, H.; An, W.; Wang, Y.; Feng, Y.; Men, Y. Synergy of a Metallic NiCo Dimer Anchored on a C2N-Graphene Matrix Promotes the Electrochemical CO2 Reduction Reaction. ACS Sustain. Chem. Eng. 2019, 7, 19113–19121. [Google Scholar] [CrossRef]
  33. Ding, C.; Feng, C.; Mei, Y.; Liu, F.; Wang, H.; Dupuis, M.; Li, C. Carbon Nitride Embedded with Transition Metals for Selective Electrocatalytic CO2 Reduction. Appl. Catal. B Environ. 2020, 268, 118391. [Google Scholar] [CrossRef]
  34. Becke, A.D.; Edgecombe, K.E. A Simple Measure of Electron Localization in Atomic and Molecular Systems. J. Chem. Phys. 1990, 92, 5397–5403. [Google Scholar] [CrossRef]
  35. Lu, T.; Chen, F.-W. Meaning and Functional Form of the Electron Localization Function. Acta Phys.-Chim. Sin. 2011, 27, 2786–2792. [Google Scholar]
  36. Pittalis, S.; Varsano, D.; Delgado, A.; Rozzi, C.A. Bonds, Lone Pairs, and Shells Probed by Means of On-Top Dynamical Correlations. Eur. Phys. J. B 2018, 91, 187. [Google Scholar] [CrossRef]
  37. Gao, Z.; Ma, H.; Yuan, S.; Ren, H.; Ge, Z.; Zhu, H.; Guo, W.; Ding, F.; Zhao, W. Benzenehexol-Based 2D MOF as High-Performance Electrocatalyst for Oxygen Reduction Reaction. Appl. Surf. Sci. 2022, 601, 154187. [Google Scholar] [CrossRef]
  38. Xing, G.; Liu, S.; Liu, J.-Y. Computational Evaluation of 2D Metal-Organic Frameworks with TMX4-Centers (X = O, S, and Se) for CO2 Electroreduction. Int. J. Hydrog. Energy 2023, 48, 3486–3494. [Google Scholar] [CrossRef]
  39. Wu, S.-Y.; Chiu, K.-Y.; Fan, C.-H.; Chen, H.-L. Electrocatalytic Carbon Dioxide Reduction on Graphene-Supported Ni Cluster and Its Hydride: Insight from First-Principles Calculations. Appl. Surf. Sci. 2023, 629, 157418. [Google Scholar] [CrossRef]
  40. Hong, X.; Chan, K.; Tsai, C.; Nørskov, J.K. How Doped MoS2 Breaks Transition-Metal Scaling Relations for CO2 Electrochemical Reduction. ACS Catal. 2016, 6, 4428–4437. [Google Scholar] [CrossRef]
  41. Chiu, K.-Y.; Fan, C.-H.; Hsu, C.-W.; Chen, H.-L. First-Principles Insight into the Mechanistic Study of Electrochemical Cyanide Reduction Reaction on Post-Transition Metal Based Single-Atom Catalysts Anchored by Phthalocyanine Nanosheets. ACS Appl. Nano Mater. 2024, 7, 9909–9924. [Google Scholar] [CrossRef]
  42. Ickmans, K.; Clarys, P.; Nijs, J.; Meeus, M.; Aerenhouts, D.; Zinzen, E.; Aelbrecht, S.; Meersdom, G.; Lambrecht, L.; Pattyn, N. Association Between Cognitive Performance, Physical Fitness, and Physical Activity Level in Women with Chronic Fatigue Syndrome. J. Rehabil. Res. Dev. 2013, 50, 795–809. [Google Scholar] [CrossRef] [PubMed]
  43. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [PubMed]
  44. Blöchl, P.E. Projector Augmented-Wave Method. Phys. Rev. B 1994, 50, 17953–17979. [Google Scholar] [CrossRef] [PubMed]
  45. Grimme, S. Semiempirical GGA-Type Density Functional Constructed with a Long-Range Dispersion Correction. J. Comput. Chem. 2006, 27, 1787–1799. [Google Scholar] [CrossRef]
Figure 1. (a) The structural diagram of TM-B@C9N4 and the transition metal (TM) atoms considered in this study. (b) The binding energy (depicted by the black line) and formation energy (depicted by the red line) of distinct TM-B@C9N4 structures. (c) Electronic localization function image of Sc-B@C9N4. (d) Band structure of Sc-B@C9N4.
Figure 1. (a) The structural diagram of TM-B@C9N4 and the transition metal (TM) atoms considered in this study. (b) The binding energy (depicted by the black line) and formation energy (depicted by the red line) of distinct TM-B@C9N4 structures. (c) Electronic localization function image of Sc-B@C9N4. (d) Band structure of Sc-B@C9N4.
Catalysts 15 00370 g001
Figure 2. (a) The adsorption configuration, bond lengths, and bond angles of CO2 on Sc-B@C9N4 and the charge density difference in the CO2-adsorbed Sc-B@C9N4. Charge depletion and accumulation are presented in cyan and yellow, respectively. The value of isosurface is 0.008 e/Å3. (b) Molecular orbitals of CO2, and PDOS of the p-orbitals of CO2 and B and the d-orbitals of TM after CO2 adsorption on the Sc-B@C9N4. (c) The PCOHP of the C-O bond after CO2 adsorption on the substrate. The dashed line was used to mark the Fermi level. The ICOHP and the length of the C-O bond are also marked in the figures (* represents substrate material).
Figure 2. (a) The adsorption configuration, bond lengths, and bond angles of CO2 on Sc-B@C9N4 and the charge density difference in the CO2-adsorbed Sc-B@C9N4. Charge depletion and accumulation are presented in cyan and yellow, respectively. The value of isosurface is 0.008 e/Å3. (b) Molecular orbitals of CO2, and PDOS of the p-orbitals of CO2 and B and the d-orbitals of TM after CO2 adsorption on the Sc-B@C9N4. (c) The PCOHP of the C-O bond after CO2 adsorption on the substrate. The dashed line was used to mark the Fermi level. The ICOHP and the length of the C-O bond are also marked in the figures (* represents substrate material).
Catalysts 15 00370 g002
Figure 3. (a) CO2RR pathway on TM-B@C9N4 for the CO product. (b) Maximum Gibbs free energy step difference for TM-B@C9N4 to reduce CO2 to CO (* represents substrate material).
Figure 3. (a) CO2RR pathway on TM-B@C9N4 for the CO product. (b) Maximum Gibbs free energy step difference for TM-B@C9N4 to reduce CO2 to CO (* represents substrate material).
Catalysts 15 00370 g003
Figure 4. The free energy distribution of CO2 to CO products under zero applied voltage (vs. RHE) in TM-B@C9N4 (* represents substrate material).
Figure 4. The free energy distribution of CO2 to CO products under zero applied voltage (vs. RHE) in TM-B@C9N4 (* represents substrate material).
Catalysts 15 00370 g004
Figure 6. (a) Structure of TM-B@C9N4-H. (b) The difference in limit potential (UL) between CO2 reduction to CO and hydrogen evolution reaction to H2 on TM-B@C9N4-H, as well as the changes in CO2 adsorption energy and H adsorption energy (* represents substrate material).
Figure 6. (a) Structure of TM-B@C9N4-H. (b) The difference in limit potential (UL) between CO2 reduction to CO and hydrogen evolution reaction to H2 on TM-B@C9N4-H, as well as the changes in CO2 adsorption energy and H adsorption energy (* represents substrate material).
Catalysts 15 00370 g006
Table 1. The Bader effective charges of TM, B, and H atoms on the surface of TM-B@C9N4(-H), where negative numbers represent electron loss and positive signs represent electron acquisition.
Table 1. The Bader effective charges of TM, B, and H atoms on the surface of TM-B@C9N4(-H), where negative numbers represent electron loss and positive signs represent electron acquisition.
SystemTi-B@C9N4(-H)V-B@C9N4(-H)Zr-B@C9N4(-H)
TM-B@C9N4TM−1.544−1.329−1.982
B−0.869−0.938−0.786
H\\\
TM-B@C9N4-HTM−1.456−1.226−1.794
B−0.811−0.776−0.771
H−0.417−0.431−0.408
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wan, R.; Zhao, B.; Li, Z. Synergistic Effect of the Heteronuclear Double Sites in C9N4 on the Electrochemical Reduction of CO2 to CO. Catalysts 2025, 15, 370. https://doi.org/10.3390/catal15040370

AMA Style

Wan R, Zhao B, Li Z. Synergistic Effect of the Heteronuclear Double Sites in C9N4 on the Electrochemical Reduction of CO2 to CO. Catalysts. 2025; 15(4):370. https://doi.org/10.3390/catal15040370

Chicago/Turabian Style

Wan, Rui, Bin Zhao, and Zhongyao Li. 2025. "Synergistic Effect of the Heteronuclear Double Sites in C9N4 on the Electrochemical Reduction of CO2 to CO" Catalysts 15, no. 4: 370. https://doi.org/10.3390/catal15040370

APA Style

Wan, R., Zhao, B., & Li, Z. (2025). Synergistic Effect of the Heteronuclear Double Sites in C9N4 on the Electrochemical Reduction of CO2 to CO. Catalysts, 15(4), 370. https://doi.org/10.3390/catal15040370

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop