ABSTRACT
Manipulating the local microenvironments of single-atom catalysts is crucial for the product selectivity of CO2 electroreduction. Although theoretical research suggests that modifying the coordination structure of isolated Ni sites can promote the reduction of CO2 to CH4, there is still no experimental evidence to date. Herein, by regulating the coordination shell of boron (B) surrounding the Ni central atom, we have achieved the transformation of the reduction product from CO to CH4. In situ techniques and density functional theory calculations reveal that B coordination in the second shell of the Ni–N–C motifs (Ni–N4–B/C) facilitates CO formation whereas incorporating B into the first shell (Ni–N3B1/C) significantly tunes the electronic structure of the Ni atoms, leading to electron delocalization, which enhances the *CO intermediate adsorption strength and makes CH4 the dominant product. This study marks the experimental realization of electrochemical CO2-to-CH4 conversion at isolated Ni sites and underscores the importance of local coordination environment regulation in steering the reaction pathways of single-atom catalysts.
Keywords: Ni single-atom catalysts, first-shell coordination, methane, local microenvironment modulation, CO2 electroreduction
Modulating the local microenvironment around single-atom sites is anticipated to steer the reaction pathway in CO2 reduction reactions. By specifically altering the coordination shell of boron around the Ni atoms, a Ni single-atom catalyst achieved a product transformation from CO to CH4, with a CH4 Faradaic efficiency of >50%.
INTRODUCTION
The continuous rise in atmospheric CO2 concentrations has captured worldwide attention for decades [1,2]. To mitigate carbon emissions and achieve carbon neutrality, the electrocatalytic CO2 reduction reaction (CO2RR) in conjunction with clean energy sources has emerged as a promising strategy for upgrading CO2 into energy-dense hydrocarbons [3–6]. However, the CO2RR process is inherently complex and involves multiple proton-coupled electron-transfer (PCET) steps as well as sophisticated intermediates, which results in a variety of products, and thus poor selectivity for a single desired product [7–9]. To address this issue, it is imperative to tune the reaction pathway carefully.
Previous research has highlighted the pivotal function of active sites and their local microenvironment modulation in steering reaction pathways [10,11]. In this context, single-atom catalysts (SACs), with ultra-high atomic utilization, well-defined geometrical configurations and tunable coordination environments, are acknowledged as exemplary model systems for adjusting reaction pathways and elucidating reaction mechanisms [12–14]. Especially, Ni SACs demonstrate the significant potential for the electrocatalysis of CO2 to CO due to their suitable adsorption and conversion for *COOH and *CO intermediates compared with other metal SACs [15–17]. Nevertheless, the elementary Ni–N4 configuration, characterized by highly symmetric geometries and electronic structures, typically hinders axial CO2 adsorption and activation, leading to high kinetic overpotentials [18,19].
Modulating the local microenvironment around the Ni sites can effectively address these challenges. Specifically, lowering the coordination number of Ni centers (NiNx, x < 4) [20–22], introducing heteroatoms (e.g. P, S, O, F) into various coordination shells [23–30] or using a combination of both strategies [31] can optimize the electronic structure of Ni atoms and alter the adsorption strength of key intermediates, ultimately affecting the reaction kinetics and selectivity for CO [32,33]. Notably, these strategies have not led to the generation of hydrocarbons such as CH4, primarily because these modifications have been tailored to enhance the adsorption of *COOH or accelerate *CO desorption. In contrast, the formation of CH4 involves a more intricate PCET mechanism. This process frequently necessitates strong *CO adsorption and adequate accessibility of protons in the reaction environment [34]. Consequently, this imposes greater demands on the design of active sites. Theoretical studies revealed that the introduction of B atoms into the first coordination shell of Ni sites can markedly enhance the electron density around the central metal, more significantly than other heteroatoms (C, N, O, S and P) [35,36]. This enhancement facilitates stronger interactions with *CO and further hydrogenation to *CHO, thereby promoting efficient CH4 production. However, this theoretical prediction has yet to be experimentally confirmed.
To elucidate the impact of local coordination environment modulation on reaction pathways, we synthesized a set of B-coordinated Ni single-atom electrocatalysts. Our findings revealed distinct product selectivity, depending on the B atom coordination shell: B positioned in the second shell facilitated only CO production whereas B in the first shell enabled CH4 formation. This regulation alters the reaction pathway, thereby enabling the product conversion from CO to CH4. The Ni–N3B1/C catalyst exhibited a peak Faradaic efficiency of 55.4% for CH4 at a total current density of 600 mA cm−2. Investigations of the mechanism and density functional theory (DFT) calculations unveiled that incorporating B into the first coordination shell effectively modulated the electronic structure of the Ni sites and elevated the d-band center. This alteration results in the delocalization of the electron distribution around the Ni atoms, thereby enhancing the Ni–C bonding interaction and promoting further hydrogenation to *CHO intermediates. Moreover, Lewis acid B sites demonstrated exceptional water-dissociation capabilities, increasing the availability of protons that were essential for CH4 production. This work is the first to achieve significant conversion of CO2 to CH4 by using Ni SACs and provides valuable insights into the potential of modifying intrinsic activity through tuning the local microenvironment of SACs.
RESULTS AND DISCUSSION
Synthesis and characterization of catalysts
The Ni single-atom catalyst simultaneously coordinated with both B and N atoms was synthesized through a refined, continuous two-step approach [37]. An illustration of the preparation procedure for the catalyst is described in Fig. 1a. Initially, a homogeneous precursor containing N, B and Ni species was prepared via solvent evaporation. Subsequently, the above precursor underwent a carbonization process at 800°C under a nitrogen atmosphere, resulting in the formation of the B, N-coordinated Ni single-atom catalyst, denoted as Ni–N3B1/C. Additionally, to shed light on the influence of the local coordination environment around the central Ni metal, additional reference materials such as Ni–N4/C and Ni–N4–B/C, were also synthesized by following a similar avenue. Inductively coupled plasma-optical emission spectroscopy analysis revealed the Ni content to be 1.24 wt% for Ni–N4/C, 1.18 wt% for Ni–N3B1/C and 1.61 wt% for Ni–N4–B/C, respectively. The similar Ni contents in the three catalysts indicate that the performance difference is determined by the local coordination environment, rather than metal loading.
Figure 1.
Synthesis and structural characterization of the Ni–N3B1/C electrocatalysts. (a) Schematic diagram of the typical preparation procedure for Ni–N3B1/C. (b) TEM image of Ni–N3B1/C. Insert in (b) displays the HRTEM image. (c) XRD patterns of Ni SACs and BN/C. (d) Aberration-corrected HAADF-STEM image of Ni–N3B1/C. (e) EELS analysis of Ni–N3B1/C. (f) The size of isolated Ni atoms based on points A1 and A2 in (d).
Scanning electron microscopy (SEM) and transmission electron microscopy (TEM) images reveal that all the as-prepared Ni SACs maintain a spherical morphology with a relatively uniform particulate size distribution (Figs S1 and S2). Further observation by using high-resolution transmission electron microscopy (HRTEM) demonstrates the absence of metal nanoparticles or clusters through the carbon substrate (Fig. 1b). The corresponding energy-dispersive X-ray spectroscopy mapping images exhibit a homogeneous dispersion of C, N, B and Ni elements in these samples (Figs S3–S5). Notably, X-ray diffraction (XRD) patterns of the catalysts, presented in Fig. 1c, exhibit only the fingerprint peaks associated with graphitic carbon. No other characteristic peaks, such as those assigned to Ni, NiOx or NiBx, are discerned, indicating the formation of highly dispersed Ni species across the entire architecture. Likewise, as illustrated in Fig. S6, the Raman spectra also identify only two vibrational bands aligned with carbon materials. To ascertain the distribution of Ni species in Ni–N3B1/C, we harnessed aberration-corrected high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM). As depicted in Fig. 1d, the carbon framework is populated by randomly dispersed scattered bright dots. These isolated bright dots, with an approximate size of 0.33 nm, represent the single Ni atoms (Fig. 1f). Electron energy loss spectroscopy (EELS), as shown in Fig. 1e, corroborates the exclusive coordination of these isolated Ni atoms adjacent to B and N atoms.
The chemical states and elemental composition were investigated by using X-ray photoelectron spectroscopy (XPS). The high-resolution Ni 2p spectra of Ni–N3B1/C reveal a slight negative shift in the binding energy of Ni 2p3/2 (855.80 eV) compared with those of Ni–N4–B/C (855.89 eV) and Ni–N4/C (855.93 eV) (Fig. 2a), indicative of the lowest oxidation state of Ni in Ni–N3B1/C. This is attributed to the incorporation of a less-electronegative heteroatom, boron, in its first coordination shell. Additional component contributions related to B–C and B–N are observed in both the Ni–N3B1/C and Ni–N4–B/C catalysts (Figs S7–S9), further demonstrating the successful integration of B atoms within the nitrogen-doped carbon substrate.
Figure 2.
Electronic structure and coordination environment analysis of Ni SACs. (a) High-resolution Ni 2p spectra of Ni–N4/C, Ni–N3B1/C and Ni–N4–B/C. (b) C K-edge XANES spectra of Ni–N4/C and Ni–N3B1/C. (c) B K-edge XANES spectra of Ni–N3B1/C. (d) Normalized Ni K-edge XANES spectra of Ni SACs and references. The inset shows the locally enlarged spectra. (e) Fourier-transformed EXAFS spectra of Ni–N3B1/C and references. (f) The corresponding EXAFS fitting curves in the R-space of Ni–N3B1/C. Wavelet transform EXAFS plots of (g) Ni–N3B1/C, (h) NiPc, (i) Ni foil and (j) NiO.
To further elucidate the electronic structure and coordination environment of Ni SACs, we employed X-ray absorption near-edge structure (XANES) spectra to probe the configurations of C, N and B elements. The C K-edge spectra (Fig. 2b) reveal three distinct peaks at 286.11 eV (Peak a), 288.85 eV (Peak b) and 293.09 eV (Peak c), which is attributed to the dipole transition of C 1s orbital electrons into π* (C=C ring), π* (C–N/B–Ni) and σ* (C–C ring) anti-bonding orbitals, respectively [38–40]. It is worth mentioning that Ni–N3B1/C presents diminished peak intensity for C=C while showing augmented peak intensity for C–N/B–Ni in comparison with Ni–N4/C. This result indicates that the incorporation of B atoms to some extent modifies the electronic structure and chemical bond within the composites, leading to the formation of more C–N/B bonds in the carbon matrix [41]. Additionally, the B K-edge spectrum for Ni–N3B1/C exhibits six characteristic peaks, assigned to B–C, B–N, B–O configurations π* state and B=N configuration σ* state, respectively (Fig. 2c) [42,43]. The N K-edge spectra of both Ni–N4/C and Ni–N3B1/C demonstrate the presence of multiple types of N species, including pyridinic-N (Peak a), pyrrolic-N (Peak b) and graphitic-N (Peak c), along with a broadened peak (Peak d) originating from the transition of N 1s orbital into σ* orbital (Fig. S10) [44,45], corresponding well with the XPS results. These observations strongly confirm the significant interactions between carbon, nitrogen and boron atoms within the catalyst architecture.
The normalized Ni K-edge XANES spectra in Fig. 2d reveal that the absorption edges of Ni–N4/C, Ni–N3B1/C and Ni–N4–B/C are situated between the Ni foil and NiO references, indicating that the average oxidation state of Ni in these three samples falls between Ni0 and Ni2+. Notably, the sequence of the absorption edges is as follows: Ni–N3B1/C < Ni–N4–B/C < Ni–N4/C, suggesting that the incorporation of B atoms results in less electron deficiency at the Ni sites compared with N atoms, which can be attributed to the weaker electronegativity of B. Furthermore, the pre-edge absorption feature (1s→3d/4p), which is highly sensitive to symmetry, provides more insight into the local geometrical structure surrounding the Ni centers. The 1s→3d transition, typically symmetry-forbidden for centrosymmetric point groups, can gain intensity in geometries allowing p–d mixing [46]. Herein, a sharper pre-edge peak at 8333.9 eV is observed for Ni–N3B1/C in comparison with Ni–N4/C, Ni–N4–B/C and NiPc, indicative of deviated quadrilateral NiN3B moieties with decreased coordination symmetry. Overall, B doping enhances the mixing of the 4p character into the 3d orbitals, leading to a stronger pre-edge absorption peak, which further supports the asymmetric incorporation of B and N atoms around the Ni centers.
The Fourier-transformed extended X-ray absorption fine structure (FT-EXAFS) spectra (Fig. 2e and Fig. S11) of Ni–N4/C, Ni–N3B1/C and Ni–N4–B/C all exhibit a dominant peak at ∼1.36 Å, corresponding to the Ni–N/Ni–B coordination in the first shell, which more closely resembles the Ni–N scattering (∼1.45 Å) observed in NiPc. No typical peak for the Ni–Ni contribution at ∼2.2 Å is noticed, confirming the absence of Ni particles or clusters and the presence of isolated Ni sites in these samples. Subsequently, quantitative EXAFS fitting was performed to investigate the fine differences in the coordination configurations of the Ni atoms on these catalysts. The detailed structural parameters are summarized in Table S1. For the Ni–N4/C catalyst, the fitting includes Ni–N and Ni–C scattering paths with a Ni–N coordination number of 3.9, indicating that the Ni atom is coordinated with four N atoms in the first shell (Fig. S12a). Meanwhile, the EXAFS fitting curves of Ni–N3B1/C unveil three scattering pathways: Ni–N, Ni–B and Ni–C, with optimal fitting results showing that the Ni atom is directly linked to three N atoms and one B atom (Fig. 2f). In contrast, for Ni–N4–B/C, along with the Ni–N path with a coordination number of 3.6, a shoulder peak at 2.72 Å is discerned, possibly ascribed to the Ni–B/Ni–C scattering in the second shell (Fig. S12b). Moreover, the k-space EXAFS fitting results are displayed in Fig. S13, providing a complementary perspective on the structural configuration. The wavelet transform (WT) of the Ni K-edge oscillations was further implemented to discriminate distinct backscattering atoms, as they can integrate information from both k-space and R-space. The Ni SACs exhibit a peak at ∼4.0 Å−1, which is entirely different from the Ni–Ni bonds (7.4 Å−1) observed in the Ni foil (Fig. 2g–j and Fig. S14), further substantiating the atomically dispersed Ni species in Ni–N4/C, Ni–N3B1/C and Ni–N4–B/C. It is noteworthy that slight WT peak shifts in the k-space of Ni–N3B1/C and Ni–N4–B/C compared with Ni–N4/C indicate discrepancies in their neighboring coordination environments.
Electrochemical CO2RR performance
To investigate the impact of different atomic coordination configurations around the Ni sites on the selectivity for CO2RR products, we conducted a series of systematic electrochemical tests. Given the poor solubility and sluggish mass transfer of CO2 in solution, the performances of catalysts were assessed by using a three-electrode flow cell reactor with 1 M KOH as the electrolyte (Fig. S15). Nuclear magnetic resonance (NMR) spectroscopy and gas chromatography (GC) were leveraged to quantify the liquid- and gas-phase products, respectively (Figs S16 and S17).
We first focused on the CO2 electrolysis for Ni–NB/C with different catalyst loadings and varying amounts of boron. Data in Figs S18 and S19 show that the CO2RR selectivity is dependent on the current density. Adding 0.1 g of boric acid culminated in a peak CH4 Faradaic efficiency (FE) of 44.1 ± 2.3%, along with a catalyst loading of 1.0 mg cm−2, which was selected for subsequent electrochemical measurements. The linear sweep voltammetry (LSV) profiles exhibit significantly enhanced current densities in a CO2 atmosphere compared with N2, signifying the active involvement of CO2 molecules in the reaction (Fig. S20). We subsequently analysed the FEs of combined products on Ni–N3B1/C and its references (Fig. 3a and Figs S21–S26). For Ni–N4/C, CO is the major product, with negligible CH4, C2H4 and formate generation. In contrast, Ni–N3B1/C demonstrates distinctively different product selectivity of CO2RR, primarily producing CH4 across the whole investigated current densities, which could be demonstrated by a remarkable GC signal for CH4 (Fig. 3b). Moreover, this catalyst achieves >10% FE for C2 products, including ethanol, acetate and ethylene, within the potential range of −1.12 to −1.35 V vs. reversible hydrogen electrode (RHE) (Fig. S25). A plausible explanation is that the occurrence of C2 product formation in parallel with CH4 might originate from *CO–CHO generation through non-adsorbed CO coupling with *CHO [47]. Intriguingly, the Ni–N4–B/C catalyst manifests a product distribution that is analogous to Ni–N4/C, maintaining an excellent CO FE of ∼100% from 100 to 400 mA cm−2. It unmasks that B doping in the second shell performs a slight influence on the electronic structure of the Ni sites, which is insufficient to alter the reaction pathway that favors the CH4 production, as previously reported [48]. These observations disclose that the improved CH4 selectivity on Ni–N3B1/C, as opposed to Ni–N4/C and Ni–N4–B/C, presumably stems from the unique coordination environment of B in the first shell (Fig. 3c). Additionally, control experiments conducted with Ni–free BN/C highlight that the CO2RR product is dominated by H2 with marginal CO and CH4 generation, suggesting that the substrates are not responsible for CH4 production (Fig. S27). Concurrently, when the electroreduction measurement of Ni–N3B1/C was conducted in N2-saturated 1 M KOH electrolyte, only H2 signals were observed, thereby substantiating that the generation of CH4 was indeed derived from CO2 molecules (Fig. S28).
Figure 3.
Electrochemical CO2-reduction performance. (a) Faradaic efficiencies for the combined products on the prepared catalysts. (b) CH4 and CO signals detected by using GC at a current density of 400 mA cm−2 on Ni–N4/C, Ni–N3B1/C and Ni–N4–B/C. (c) Faradaic efficiencies for CH4 under various applied potentials. (d) In situ DEMS measurement for CH4 production. (e) Partial current densities for CH4 at different applied potentials. (f) In situ Raman spectra of Ni–N3B1/C at various potentials, recording the peak area change of HCO3− and CO32−. (g) pH values calculated from in situ Raman spectra, as a function of potential. (h) Long-term stability test for converting CO2 to CH4 at a current density of 300 mA cm−2 over Ni–N3B1/C. Inset shows the schematic illustration of the flow cell. The error bars in (a, c, e) are generated based on the mean ± standard deviation of three independent measurements.
In situ differential electrochemical mass spectrometry (DEMS) was utilized to directly validate the CH4 formation. As illustrated in Fig. 3d and Fig. S29, Ni–N3B1/C performs lower overpotential and higher signal intensity for CO2-to-CH4 conversion relative to its counterparts, demonstrating that the first-shell B dopants indeed motivate the intrinsic activity for electrochemical CO2 reduction towards CH4. Data in Fig. 3e reveal that Ni–N3B1/C achieves a maximum partial current density of 242.3 mA cm−2 for CH4, whilst the value for Ni–N4–B/C is exclusively 58.9 mA cm−2 under the same conditions. In stark contrast, Ni–N4/C presents a negligible CH4 current density over the entire potential window. Moreover, the partial current densities of H2 and CO on these catalysts are also detailed in Fig. S30. Notably, under anion-exchange membrane conditions, the FE for CH4 further increased to 55.4 ± 0.5% with a partial current density of 332.4 mA cm−2 (Fig. S31), surpassing massive state-of-the-art CO2RR electrocatalysts (Fig. S32 and Table S2). Subsequently, we employed in situ Raman spectroscopy to monitor changes in the local pH during the reaction by analysing the peak area ratio of bicarbonate to carbonate. As depicted in Fig. 3f and g, and Fig. S33, a notable decrease in the local pH is indeed observed. This phenomenon can be well explained by the pH dependency of the hydrogenation process from *CO to *CHO species. A reduction in the pH increases the availability of protons, thereby lowering the activation energy barrier for *CHO production and eventually facilitating the formation of methane [49].
To acquire a high FECH4 of >50%, we initially executed a long-term galvanostatic test at a current density of 500 mA cm−2 using the Ni–N3B1/C catalyst. Unfortunately, we noticed a significant decline in FECH4 owing to the loss of hydrophobicity in the gas diffusion electrode (GDE), accompanied by rapid hydrogen evolution after only 1 h (Fig. S34). Afterward, we adjusted the applied current density to 300 mA cm−2 at the expense of the FECH4 value. Remarkably, our Ni–N3B1/C catalyst exhibited superior stability, as evidenced by the absence of noticeable fluctuations in both CH4 FE and overpotential over 34 operating hours (Fig. 3h). A suite of structural and morphological analyses also revealed that, following electrolysis, Ni–N3B1/C still retained its monatomic dispersion (Fig. S35).
Mechanism investigation
To better understand the catalytic mechanism of the promotive effect of B doping, in situ attenuated total reflection surface-enhanced infrared absorption spectroscopy (ATR-SEIRAS) was employed to monitor the reactive species and key intermediates over the catalyst surface. As depicted in Fig. 4a, for the Ni–N3B1/C catalyst, the peak around ∼1637 cm−1 corresponds to bending vibrations of the interfacial H2O, suggesting the participation of water in the electrolysis process [18,50]. With the decrease in applied potentials, a gradually enhanced band at ∼1218 cm−1 can be ascribed to the OH deformation of *COOH, which is generally accepted as the pivotal intermediate for CO2RR to CO and CH4 [51,52]. Furthermore, the peak centered at ∼1475 cm−1, with an obvious Stark effect, is possibly regarded as the CO32− species [53,54]. Similar species are also detected in Ni–N4/C, but with lower characteristic intensities (Fig. 4c). Significantly, however, several new peaks associated with *CH3O (∼1408 cm−1) [55–57] and CHx (∼2866 and ∼2953 cm−1) [58], which are the crucial intermediates for methane formation, are distinctly observed only at −0.5 V vs. RHE over Ni–N3B1/C (Fig. 4b and Fig. S36a). Conversely, for the Ni–N4/C catalyst, a weak *CH3O signal is not shown until −1.0 V vs. RHE, with no significant feature of CHx across the entire potential region (Fig. 4d and Fig. S36b). This phenomenon corresponds well with our experimental results, indicating that the introduction of B contributes to the formation of hydrogenated intermediates, thereby steering different product distributions.
Figure 4.
Investigation of key intermediates. In situ ATR-SEIRAS spectra recorded at various applied potentials for (a, b) Ni–N3B1/C and (c, d) Ni–N4/C in CO2-saturated 0.1 M KHCO3 electrolyte. (e) Electrochemical CO-stripping curves for Ni–N3B1/C and Ni–N4/C in 0.1 M KHCO3 electrolyte. The solid and dashed curves correspond to the first (presence of CO adsorbed on the catalyst) and second (absence of CO absorbed on the catalyst) scans, respectively. (f) Single oxidative LSV scans in N2-saturated 0.1 M KOH solution for Ni–N3B1/C and Ni–N4/C.
To examine the influence of this modification on CO-adsorption capability, we conducted electrochemical CO-stripping experiments (Fig. 4e) [59,60]. The Ni–N3B1/C catalyst exhibits a more pronounced and more positive CO-stripping peak than the Ni–N4/C catalyst, indicating a stronger CO-binding ability, which is advantageous for the subsequent hydrogenation of CO to generate methane. On the other hand, we performed OH−-adsorption measurements to further evaluate the binding affinity of *CO2− on catalysts. As shown in Fig. 4f, Ni–N3B1/C exhibits a more positive potential for surface OH− adsorption compared with Ni–N4/C, implying a more effective stabilization of *CO2− on Ni–N3B1/C. Altogether, these findings prove that Ni–N3B1/C is capable of boosting CO2 molecule activation and enhancing CO adsorption, ultimately accelerating the conversion of CO2 to CH4.
Previous investigations into Ni SACs have predominantly resulted in CO formation, while the subsequent hydrogenation to methane remains challenging. This is largely attributed to the relatively weak adsorption energy of *CO intermediates on Ni sites. To enhance methane production, it is imperative to improve the adsorption of *CO intermediates to enable further hydrogenation to *CHO. This process heavily relies on optimizing the electronic structure of the isolated Ni sites. Utilizing DFT calculations, we explored the influence of the coordination environment around the Ni center on the catalytic mechanism for CO2 electroreduction to CH4. Based on the EXAFS fitting results, three models, namely Ni–N4–C, Ni–N4–B–C and Ni–N3B1–C, were constructed to represent the Ni–N4/C, Ni–N4–B/C and Ni–N3B1/C catalysts, respectively. All computational structural models are illustrated in Figs S37–S39. The projected density of states (PDOS) analysis (Fig. 5a) indicates that the incorporation of B into the first shell significantly elevates the d-band center of Ni atoms, resulting in electron delocalization, which enhances hybridization between the d orbital and the anti-bonding 2π* orbital of CO (known as d→2π* backdonation), facilitating more profound CO2 reduction [61–63]. Subsequently, we adopted a computational hydrogen electrode model to investigate the free energy diagram for the conversion of CO2 to *CHO—a crucial intermediate for CH4 production (Fig. 5b). The rate-determining step for all these catalysts was identified as the conversion from *CO2 to *COOH and this step was favorable on Ni–N3B1/C. After *CO formation, *CO will hydrogenate to *CHO, which is exothermic at the Ni–N3B1/C site, while endothermic at the Ni–N4/C and Ni–N4–B/C sites, indicating a preference for CH4 production instead of CO on the Ni–N3B1/C catalyst. This observation is attributed to the enhanced *CO-adsorption energy (Fig. S40). Moreover, the free energies for each intermediate on Ni–N3B1/C were found to be lower than those on Ni–N4/C and Ni–N4–B/C, further emphasizing the superiority of B coordination in the first shell for promoting CO2 activation and intermediate stability.
Figure 5.
DFT calculations. (a) Projected density of states (PDOS) and (b) Gibbs free energy profiles of CO2RR to CH4 pathways on Ni–N4/C, Ni–N4–B/C and Ni–N3B1/C. The d-band centers were calculated as −0.712 eV for Ni–N3B1/C, −1.661 eV for Ni–N4–B/C and −1.665 eV for Ni–N4/C. PDOS for Ni 3d orbital and 2p orbital of adsorbed CO as well as crystal orbital Hamilton population (COHP) analysis of the Ni–C bond for (c) Ni–N4/C, (d) Ni–N4–B/C and (e) Ni–N3B1/C. Two-dimensional differential charge density cross section of (f) Ni–N4/C, (g) Ni–N4–B/C and (h) Ni–N3B1/C. (i) Energy barrier profiles for H2O dissociation process on Ni–N4/C and Ni–N3B1/C. (j) KIE of CO2RR to CH4 on Ni–N3B1/C. (k) Calculated Cφ of Ni–N4/C and Ni–N3B1/C at various potentials.
Given the pivotal role of *CO adsorption in regulating product selectivity, we further employed PDOS and crystal orbital Hamilton population (COHP) analyses to elucidate the interactions between *CO and catalysts. Data in Fig. 5c–e illustrate an enhanced orbital overlap between Ni 3d for Ni–N3B1/C and C 2p for the adsorbed *CO compared with Ni–N4/C and Ni–N4–B/C. COHP analysis reveals a substantial anti-bonding contribution near the Fermi energy level for Ni–N4/C and Ni–N4–B/C whereas Ni–N3B1/C shows minimal anti-bonding contributions. Further integration of the COHP confirms a more robust Ni–C bond interaction on Ni–N3B1/C, which is advantageous for *CO adsorption and subsequent hydrogenation to CH4. This conclusion is further supported by visualized differential charge distribution (Fig. S41). A more pronounced electron transfer and shorter Ni–C bond length (Table S3) between the Ni center and *CO intermediates are observed in the presence of the first-shell B atom, suggesting that the introduction of B can improve d→2π* backdonation and thus stabilize the *CO species (Fig. 5f–h). To quantify electron transfer during CO adsorption, we conducted a Mulliken charge population analysis. Intriguingly, B also functioned as an additional electron donor for *CO adsorption, alongside Ni sites (Table S4), suggesting that the adsorption of intermediates resulted from the synergistic effect of the Ni and B sites.
For the formation of CH4, enhancing CO adsorption is crucial, while the availability of protons is equally essential. Under alkaline conditions, water molecules serve as the sole proton source. Improving water dissociation can supply more available protons for methane production. In this study, we selected the free energy change (ΔG) for H2O dissociation as a crucial descriptor to assess the capability of proton supply in the reaction. As shown in Fig. 5i, the ΔG for H2O dissociation at both the Ni (0.99 eV) and B (0.57 eV) sites on Ni–N3B1/C was significantly lower than that on Ni–N4/C (3.33 eV), demonstrating that B doping accelerates water dissociation and improves the proton-feeding ability, facilitating the generation of key intermediates such as *COOH and *CHO. This finding aligns with prior research indicating that the Lewis acid B sites, owing to their oxygenophilic properties, favor the adsorption of water molecules and the cleavage of H–OH bonds [26,64]. Subsequently, the pivotal role of B atoms in water activation and protonation processes was further corroborated through kinetic isotope effect (KIE) measurements of H/D on Ni–N3B1/C (Fig. 5j and Fig. S42). The measured KIE values were <1.0 at both −1.2 V and −1.4 V vs. RHE, indicating that hydrogenation is not the rate-determining step on the Ni–N3B1/C catalyst [51,65,66]. In addition, we carried out in situ electrochemical impedance spectroscopy (EIS) to examine the availability of adsorped hydrogen (H*) on the catalyst surface. A double-parallel equivalent circuit model was utilized to simulate the Nyquist plots and adsorption pseudocapacitance (Cφ) was applied to quantify the H* availability (Fig. S43) [67,68]. Compared with Ni–N4/C, Ni–N3B1/C exhibited greater H* availability, which could provide sufficient H* for the hydrogenation process (Fig. 5k). These comprehensive analyses collectively indicate that the generation of CH4 on Ni–N3B1/C benefits from both the enhanced *CO adsorption and accelerated water-dissociation kinetics.
CONCLUSION
In conclusion, we developed and synthesized a Ni single-atom catalyst coordinated with boron and nitrogen in the first shell. Notably, the presence of boron in the first shell facilitated the generation of CH4 whereas its absence or presence in the secondary shell led exclusively to CO formation. This observation underscores the crucial role of the coordination environment at the Ni sites in steering the reaction pathway. Combined with in situ characterization and DFT calculations, it was demonstrated that the introduction of B into the first shell significantly elevated the d-band center of Ni, induced its electron delocalization and enhanced CO adsorption. Furthermore, boron served as an active site for water dissociation, thereby accelerating the kinetics of proton-coupled electron transfer. Taken together, the Ni–N3B1/C catalyst successfully achieved > 50% CH4 FE on isolated Ni sites for the first time. This study highlights the significant impact of local chemical microenvironment regulation on reaction pathways, while also providing profound insights into the design of non-copper-based electrocatalysts for promoting the generation of hydrocarbons.
Supplementary Material
ACKNOWLEDGEMENTS
We also acknowledge the Instrumental Analysis Centre of Shenzhen University for performing TEM and testing 1H-NMR.
Contributor Information
Yan Kong, College of Chemistry and Environmental Engineering, Shenzhen University, Shenzhen 518060, China; Department of Chemical Physics, University of Science and Technology of China, Hefei 23002, China.
Xinmei Jia, College of Chemistry and Environmental Engineering, Shenzhen University, Shenzhen 518060, China; Department of Chemical Physics, University of Science and Technology of China, Hefei 23002, China.
Xiaoyan Chai, College of Chemistry and Environmental Engineering, Shenzhen University, Shenzhen 518060, China.
Zhi Chen, College of Chemistry and Environmental Engineering, Shenzhen University, Shenzhen 518060, China.
Chunyan Shang, College of Chemistry and Environmental Engineering, Shenzhen University, Shenzhen 518060, China.
Xingxing Jiang, College of Chemistry and Environmental Engineering, Shenzhen University, Shenzhen 518060, China.
Huizhu Cai, College of Chemistry and Environmental Engineering, Shenzhen University, Shenzhen 518060, China.
Lingyan Jing, College of Chemistry and Environmental Engineering, Shenzhen University, Shenzhen 518060, China.
Qi Hu, College of Chemistry and Environmental Engineering, Shenzhen University, Shenzhen 518060, China.
Hengpan Yang, College of Chemistry and Environmental Engineering, Shenzhen University, Shenzhen 518060, China.
Xue Zhang, College of Chemistry and Environmental Engineering, Shenzhen University, Shenzhen 518060, China.
Chuanxin He, College of Chemistry and Environmental Engineering, Shenzhen University, Shenzhen 518060, China.
FUNDING
This work was supported by the National Natural Science Foundation of China (U21A20312 and 22472105) and the Guangdong Basic and Applied Basic Research Foundation (2022B1515120084).
AUTHOR CONTRIBUTIONS
C. X. H., X. Z., H. P. Y. and Q. H. supervised the project. Y. K. conceived the concept, designed and carried out the experiments and wrote the manuscript. X. M. J. assisted with the electrochemical measurements. X. Y. C. and Z. C. provided technical support in material characterization. X. Z. assisted in the analysis of the DFT calculations. X. X. J., C. Y. S., H. Z. C. and L. Y. J. aided in writing and data analysis.
Conflict of interest statement. None declared.
REFERENCES
- 1. Chu S, Cui Y, Liu N. The path towards sustainable energy. Nat Mater 2017; 16: 16–22. 10.1038/nmat4834 [DOI] [PubMed] [Google Scholar]
- 2. Olah GA, Prakash GKS, Goeppert A. Anthropogenic chemical carbon cycle for a sustainable future. J Am Chem Soc 2011; 133: 12881–98. 10.1021/ja202642y [DOI] [PubMed] [Google Scholar]
- 3. Ross MB, De Luna P, Li Y et al. Designing materials for electrochemical carbon dioxide recycling. Nat Catal 2019; 2: 648–58. 10.1038/s41929-019-0306-7 [DOI] [Google Scholar]
- 4. Whipple DT, Kenis PJA. Prospects of CO2 utilization via direct heterogeneous electrochemical reduction. J Phys Chem Lett 2010; 1: 3451–8. 10.1021/jz1012627 [DOI] [Google Scholar]
- 5. Xu S, Carter EA. Theoretical insights into heterogeneous (photo)electrochemical CO2 reduction. Chem Rev 2019; 119: 6631–69. 10.1021/acs.chemrev.8b00481 [DOI] [PubMed] [Google Scholar]
- 6. Jiang Y, Huang L, Chen C et al. Catalyst-electrolyte interface engineering propels progress in acidic CO2 electroreduction. Energy Environ Sci 2025; 18: 2025–49. 10.1039/D4EE05715E [DOI] [Google Scholar]
- 7. Hori Y, Kikuchi K, Suzuki S. Production of CO and CH4 in electrochemical reduction of CO2 at metal electrodes in aqueous hydrogencarbonate solution. Chem Lett 1985; 14: 1695–8. 10.1246/cl.1985.1695 [DOI] [Google Scholar]
- 8. Zhang J, Ding J, Liu Y et al. Molecular tuning for electrochemical CO2 reduction. Joule 2023; 7: 1700–44. 10.1016/j.joule.2023.07.010 [DOI] [Google Scholar]
- 9. Kortlever R, Shen J, Schouten KJP et al. Catalysts and reaction pathways for the electrochemical reduction of carbon dioxide. J Phys Chem Lett 2015; 6: 4073–82. 10.1021/acs.jpclett.5b01559 [DOI] [PubMed] [Google Scholar]
- 10. Huang H, Jung H, Li S et al. Activation of inert copper for significantly enhanced hydrogen evolution behaviors by trace ruthenium doping. Nano Energy 2022; 92: 106763. 10.1016/j.nanoen.2021.106763 [DOI] [Google Scholar]
- 11. Zhang E, Tao L, An J et al. Engineering the local atomic environments of indium single-atom catalysts for efficient electrochemical production of hydrogen peroxide. Angew Chem Int Ed 2022; 61: e202117347. 10.1002/anie.202117347 [DOI] [PubMed] [Google Scholar]
- 12. Fei H, Dong J, Feng Y et al. General synthesis and definitive structural identification of MN4C4 single-atom catalysts with tunable electrocatalytic activities. Nat Catal 2018; 1: 63–72. 10.1038/s41929-017-0008-y [DOI] [Google Scholar]
- 13. Yang X-F, Wang A, Qiao B et al. Single-atom catalysts: a new frontier in heterogeneous catalysis. Acc Chem Res 2013; 46: 1740–8. 10.1021/ar300361m [DOI] [PubMed] [Google Scholar]
- 14. Wang Y, Wang D, Li Y. Rational design of single-atom site electrocatalysts: from theoretical understandings to practical applications. Adv Mater 2021; 33: 2008151. 10.1002/adma.202008151 [DOI] [PubMed] [Google Scholar]
- 15. Li Y, Adli NM, Shan W et al. Atomically dispersed single Ni site catalysts for high-efficiency CO2 electroreduction at industrial-level current densities. Energy Environ Sci 2022; 15: 2108–19. 10.1039/D2EE00318J [DOI] [Google Scholar]
- 16. Zheng T, Jiang K, Ta N et al. Large-scale and highly selective CO2 electrocatalytic reduction on nickel single-atom catalyst. Joule 2019; 3: 265–78. 10.1016/j.joule.2018.10.015 [DOI] [Google Scholar]
- 17. Hao Q, Zhong H, Wang J et al. Nickel dual-atom sites for electrochemical carbon dioxide reduction. Nat Synth 2022; 1: 719–28. 10.1038/s44160-022-00138-w [DOI] [Google Scholar]
- 18. Chen J, Li Z, Wang X et al. Promoting CO2 electroreduction kinetics on atomically dispersed monovalent ZnI sites by rationally engineering proton-feeding centers. Angew Chem Int Ed 2022; 61: e202111683. 10.1002/anie.202111683 [DOI] [PubMed] [Google Scholar]
- 19. Ren W, Tan X, Yang W et al. Isolated diatomic Ni–Fe metal-nitrogen sites for synergistic electroreduction of CO2. Angew Chem Int Ed 2019; 58: 6972–6. 10.1002/anie.201901575 [DOI] [PubMed] [Google Scholar]
- 20. Zhou Y, Zhou Q, Liu H et al. Asymmetric dinitrogen-coordinated nickel single-atomic sites for efficient CO2 electroreduction. Nat Commun 2023; 14: 3776. 10.1038/s41467-023-39505-2 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21. Zhang Y, Jiao L, Yang W et al. Rational fabrication of low-coordinate single-atom Ni electrocatalysts by MOFs for highly selective CO2 reduction. Angew Chem Int Ed 2021; 60: 7607–11. 10.1002/anie.202016219 [DOI] [PubMed] [Google Scholar]
- 22. Zhang W, Liu D, Liu T et al. Coordinately unsaturated nickel single atom electrocatalyst for efficient CO2 conversion. Nano Res 2023; 16: 10873–80. 10.1007/s12274-023-5949-7 [DOI] [Google Scholar]
- 23. Chen S, Li X, Kao C-W et al. Unveiling the proton-feeding effect in sulfur-doped Fe–N–C single-atom catalyst for enhanced CO2 electroreduction. Angew Chem Int Ed 2022; 61: e202206233. 10.1002/anie.202206233 [DOI] [PubMed] [Google Scholar]
- 24. Sun X, Wang L, Lan X et al. Positively charged nickel-sulfur dual sites for efficient CO2 electroreduction reaction. Appl Catal, B 2024; 342: 123389. 10.1016/j.apcatb.2023.123389 [DOI] [Google Scholar]
- 25. Chen Z, Wang C, Zhong X et al. Achieving efficient CO2 electrolysis to CO by local coordination manipulation of nickel single-atom catalysts. Nano Lett 2023; 23: 7046–53. 10.1021/acs.nanolett.3c01808 [DOI] [PubMed] [Google Scholar]
- 26. Zhang Q, Tsai HJ, Li F et al. Boosting the proton-coupled electron transfer via Fe-P atomic pair for enhanced electrochemical CO2 reduction. Angew Chem Int Ed 2023; 62: e202311550. 10.1002/anie.202311550 [DOI] [PubMed] [Google Scholar]
- 27. Wang Z, Chen Z, Du X et al. Modulating the electronic properties of single Ni atom catalyst via first-shell coordination engineering to boost electrocatalytic flue gas CO2 reduction. Adv Funct Mater 2025; 35: 2420994. 10.1002/adfm.202420994 [DOI] [Google Scholar]
- 28. Miao K, Qin J, Lai S et al. Spin regulation of nickel single atom catalyst via axial phosphor-coordination achieves near unity CO selectivity in electrochemical CO2 reduction. Adv Funct Mater 2025; 35: 2419989. 10.1002/adfm.202419989 [DOI] [Google Scholar]
- 29. Qu M, Chen Z, Sun Z et al. Rational design of asymmetric atomic Ni-P1N3 active sites for promoting electrochemical CO2 reduction. Nano Res 2023; 16: 2170–6. 10.1007/s12274-022-4969-z [DOI] [Google Scholar]
- 30. Wang Y, Zhu P, Wang R et al. Fluorine-tuned carbon-based nickel single-atom catalysts for scalable and highly efficient CO2 electrocatalytic reduction. ACS Nano 2024; 18: 26751–8. 10.1021/acsnano.4c06923 [DOI] [PubMed] [Google Scholar]
- 31. Jia C, Tan X, Zhao Y et al. Sulfur-dopant-promoted electroreduction of CO2 over coordinatively unsaturated Ni-N2 moieties. Angew Chem Int Ed 2021; 60: 23342–8. 10.1002/anie.202109373 [DOI] [PubMed] [Google Scholar]
- 32. Gao Y, Liu B, Wang D. Microenvironment engineering of single/dual-atom catalysts for electrocatalytic application. Adv Mater 2023; 35: 2209654. 10.1002/adma.202209654 [DOI] [PubMed] [Google Scholar]
- 33. Liu P, Zheng N. Coordination chemistry of atomically dispersed catalysts. Natl Sci Rev 2018; 5: 636–8. 10.1093/nsr/nwy051 [DOI] [Google Scholar]
- 34. Peterson AA, Nørskov JK. Activity descriptors for CO2 electroreduction to methane on transition-metal catalysts. J Phys Chem Lett 2012; 3: 251–8. 10.1021/jz201461p [DOI] [Google Scholar]
- 35. Zhang Y, Liu T, Wang X et al. Dual-atom metal and nonmetal site catalyst on a single nickel atom supported on a hybridized BCN nanosheet for electrochemical CO2 reduction to methane: combining high activity and selectivity. ACS Appl Mater Interfaces 2022; 14: 9073–83. 10.1021/acsami.1c22761 [DOI] [PubMed] [Google Scholar]
- 36. Fu Z, Li Q, Bai X et al. Promoting the conversion of CO2 to CH4 via synergistic dual active sites. Nanoscale 2021; 13: 12233–41. 10.1039/D1NR02582A [DOI] [PubMed] [Google Scholar]
- 37. Min Y, Zhou X, Chen J-J et al. Integrating single-cobalt-site and electric field of boron nitride in dechlorination electrocatalysts by bioinspired design. Nat Commun 2021; 12: 303. 10.1038/s41467-020-20619-w [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38. Chen P, Zhang N, Wang S et al. Interfacial engineering of cobalt sulfide/graphene hybrids for highly efficient ammonia electrosynthesis. Proc Natl Acad Sci USA 2019; 116: 6635–40. 10.1073/pnas.1817881116 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39. Wan J, Zhao Z, Shang H et al. In situ phosphatizing of triphenylphosphine encapsulated within metal-organic frameworks to design atomic Co1-P1N3 interfacial structure for promoting catalytic performance. J Am Chem Soc 2020; 142: 8431–9. 10.1021/jacs.0c02229 [DOI] [PubMed] [Google Scholar]
- 40. Bruhwiler PA, Maxwell AJ, Puglia C et al. π* and σ* excitions in C-1s absorption of graphite. Phys Rev Lett 1995; 74: 614–7. 10.1103/PhysRevLett.74.614 [DOI] [PubMed] [Google Scholar]
- 41. Tong Y, Chen P, Zhou T et al. A bifunctional hybrid electrocatalyst for oxygen reduction and evolution: cobalt oxide nanoparticles strongly coupled to B,N-decorated graphene. Angew Chem Int Ed 2017; 56: 7121–5. 10.1002/anie.201702430 [DOI] [PubMed] [Google Scholar]
- 42. Zhang X, Yan P, Xu J et al. Methanol conversion on borocarbonitride catalysts: identification and quantification of active sites. Sci Adv 2020; 6: eaba5778. 10.1126/sciadv.aba5778 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43. Chen S, Chen Z, Siahrostami S et al. Designing boron nitride islands in carbon materials for efficient electrochemical synthesis of hydrogen peroxide. J Am Chem Soc 2018; 140: 7851–9. 10.1021/jacs.8b02798 [DOI] [PubMed] [Google Scholar]
- 44. Mei B, Liu C, Sun F et al. Unraveling the potential-dependent volcanic selectivity changes of an atomically dispersed Ni catalyst during CO2 reduction. ACS Catal 2022; 12: 8676–86. 10.1021/acscatal.2c01885 [DOI] [Google Scholar]
- 45. Zhao Y, Yang N, Yao H et al. Stereodefined codoping of sp-N and S atoms in few-layer graphdiyne for oxygen evolution reaction. J Am Chem Soc 2019; 141: 7240–4. 10.1021/jacs.8b13695 [DOI] [PubMed] [Google Scholar]
- 46. Colpas GJ, Maroney MJ, Bagyinka C et al. X-ray spectroscopic studies of nickel complexes, with application to the structure of nickel sites in hydrogenases. Inorg Chem 1991; 30: 920–8. 10.1021/ic00005a010 [DOI] [Google Scholar]
- 47. Cheng T, Xiao H, Goddard WA. Full atomistic reaction mechanism with kinetics for CO reduction on Cu(100) from ab initio molecular dynamics free-energy calculations at 298 K. Proc Natl Acad Sci USA 2017; 114: 1795–800. 10.1073/pnas.1612106114 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48. Gu X, Jiao Y, Wei B et al. Boron bridged NiN4B2Cx single-atom catalyst for superior electrochemical CO2 reduction. Mater Today 2022; 54: 63–71. 10.1016/j.mattod.2022.02.008 [DOI] [Google Scholar]
- 49. Zhu J, Lv L, Zaman S et al. Advances and challenges in single-site catalysts towards electrochemical CO2 methanation. Energy Environ Sci 2023; 16: 4812–33. 10.1039/D3EE02196C [DOI] [Google Scholar]
- 50. Jiang Y, Li H, Chen C et al. Dynamic Cu0/Cu+ interface promotes acidic CO2 electroreduction. ACS Catal 2024; 14: 8310–6. 10.1021/acscatal.4c01516 [DOI] [Google Scholar]
- 51. Chen S, Zhang Z, Jiang W et al. Engineering water molecules activation center on multisite electrocatalysts for enhanced CO2 methanation. J Am Chem Soc 2022; 144: 12807–15. 10.1021/jacs.2c03875 [DOI] [PubMed] [Google Scholar]
- 52. Chen S, Li W, Jiang W et al. MOF encapsulating N-heterocyclic carbene-ligated copper single-atom site catalyst towards efficient methane electrosynthesis. Angew Chem Int Ed 2022; 61: e202114450. 10.1002/anie.202114450 [DOI] [PubMed] [Google Scholar]
- 53. Delmo EP, Wang Y, Song Y et al. In situ infrared spectroscopic evidence of enhanced electrochemical CO2 reduction and C–C coupling on oxide-derived copper. J Am Chem Soc 2024; 146: 1935–45. 10.1021/jacs.3c08927 [DOI] [PubMed] [Google Scholar]
- 54. Zhang L, Feng J, Liu S et al. Atomically dispersed Ni-Cu catalysts for pH-universal CO2 electroreduction. Adv Mater 2023; 35: 2209590. 10.1002/adma.202209590 [DOI] [PubMed] [Google Scholar]
- 55. Zhao J, Zhang P, Yuan T et al. Modulation of *CHxO adsorption to facilitate electrocatalytic reduction of CO2 to CH4 over Cu-based catalysts. J Am Chem Soc 2023; 145: 6622–7. 10.1021/jacs.2c12006 [DOI] [PubMed] [Google Scholar]
- 56. Zhou X, Shan J, Chen L et al. Stabilizing Cu2+ ions by solid solutions to promote CO2 electroreduction to methane. J Am Chem Soc 2022; 144: 2079–84. 10.1021/jacs.1c12212 [DOI] [PubMed] [Google Scholar]
- 57. Chen X, Jia S, Chen C et al. Highly stable layered coordination polymer electrocatalyst toward efficient CO2-to-CH4 conversion. Adv Mater 2024; 36: 2310273. 10.1002/adma.202310273 [DOI] [PubMed] [Google Scholar]
- 58. Song X, Xiong W, He H et al. Boosting CO2 electrocatalytic reduction to ethylene via hydrogen-assisted C–C coupling on Cu2O catalysts modified with Pd nanoparticles. Nano Energy 2024; 122: 109275. 10.1016/j.nanoen.2024.109275 [DOI] [Google Scholar]
- 59. Guo W, Liu S, Tan X et al. Highly efficient CO2 electroreduction to methanol through atomically dispersed Sn coupled with defective CuO catalysts. Angew Chem Int Ed 2021; 60: 21979–87. 10.1002/anie.202108635 [DOI] [PubMed] [Google Scholar]
- 60. Feng J, Zhang L, Liu S et al. Modulating adsorbed hydrogen drives electrochemical CO2-to-C2 products. Nat Commun 2023; 14: 4615. 10.1038/s41467-023-40412-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61. Gunathunge CM, Li J, Li X et al. Surface-adsorbed CO as an infrared probe of electrocatalytic interfaces. ACS Catal 2020; 10: 11700–11. 10.1021/acscatal.0c03316 [DOI] [Google Scholar]
- 62. Blyholder G. Molecular orbital view of chemisorbed carbon monoxide. J Phys Chem 1964; 68: 2772–7. 10.1021/j100792a006 [DOI] [Google Scholar]
- 63. Föhlisch A, Nyberg M, Hasselström J et al. How carbon monoxide adsorbs in different sites. Phys Rev Lett 2000; 85: 3309–12. 10.1103/PhysRevLett.85.3309 [DOI] [PubMed] [Google Scholar]
- 64. Qin Q, Jang H, Jiang X et al. Constructing interfacial oxygen vacancy and ruthenium Lewis acid-base pairs to boost the alkaline hydrogen evolution reaction kinetics. Angew Chem Int Ed 2024; 63: e202317622. 10.1002/anie.202317622 [DOI] [PubMed] [Google Scholar]
- 65. Ma W, Xie S, Zhang X et al. Promoting electrocatalytic CO2 reduction to formate via sulfur-boosting water activation on indium surfaces. Nat Commun 2019; 10: 892. 10.1038/s41467-019-08805-x [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66. Wang X, Sang X, Dong C-L et al. Proton capture strategy for enhancing electrochemical CO2 reduction on atomically dispersed metal–nitrogen active sites. Angew Chem Int Ed 2021; 60: 11959–65. 10.1002/anie.202100011 [DOI] [PubMed] [Google Scholar]
- 67. Li J, Liu H, Gou W et al. Ethylene-glycol ligand environment facilitates highly efficient hydrogen evolution of Pt/CoP through proton concentration and hydrogen spillover. Energy Environ Sci 2019; 12: 2298–304. 10.1039/C9EE00752K [DOI] [Google Scholar]
- 68. Chen S, Wang S, Hao P et al. N,O-C nanocage-mediated high-efficient hydrogen evolution reaction on IrNi@N,O-C electrocatalyst. Appl Catal, B 2022; 304: 120996. 10.1016/j.apcatb.2021.120996 [DOI] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.





