Theoretical prediction of Ni@NiNCM model. We first performed computational studies to predict the rationally designed Ni@NiNxCM architecture for boosting the catalytic activities. As previously reported5,15, a NiN4CM catalyst exhibited excellent electrocatalytic performance for CO2-to-CO conversion, originating from the NiN4CM’s d-states that played an important role in the CO2 activation and protonation processes16,17. As such, one of the decisive steps to improve the intrinsic activity for boosting the ECR performance is to rationally modulate the d-states of the NiN4CM catalyst. By analog, therefore, we optimized the geometric structure of Ni@NiN4CM through the density of state calculations by CASTEP (Figs. 1a-1c). Figures 1d and 1e showed the partial density of states (PDOS) for Ni@NiN4CM without and with adsorbed CO2. As seen in Fig. 1d, the Ni@NiN4CM model shows a strong peak of the highest occupied molecular orbital (HOMO) comprised of strengthened d-states from the Ni atoms18. Meanwhile, the states around the Fermi level indicate that the Ni@NiN4CM model has a remarkable electron transport capability. These are two key factors for regulating the catalytic activities. As shown in Fig. 1e, the HOMO peak of Ni@NiN4CM downshifted only slightly compared with that in Fig. 1d, indicating that the active center and its electronic property could be well retained to show an excellent activity during the catalytic reactions.
Structural characterization of Ni@NiNCM. As schematically shown in Fig. 2a, the Ni@NiNCM was synthesized through N2-protected carbonization of a precursor that was obtained via a simple rotary evaporation (42 oC) of an ethanol mixture solution of nickel nitrate (Ni(NO3)2·6H2O) and o-phenylenediamine (oPD) (WNi/WoPD is 1/10). Possible effects of the carbonization temperature on the ECR performance of Ni@NiNCM were investigated in details over 700–1,000 oC, and the optimum temperature was found to be 900 oC (Supplementary Figs. 1–3).
Transmission electron microscopy (TEM) image of the as-prepared Ni@NiNCM shown in Fig. 2b displays a typical layered morphology characteristic of carbon nanosheets with the interspersed Ni NPs of ~ 15 nm in diameter. The corresponding high-resolution TEM (HRTEM) image shows an interplanar spacing of 2.0 Å, in consistent well with the (111) crystal plane of metallic Ni NPs (Fig. 2c). However, X-ray diffraction (XRD) pattern of Ni@NiNCM only exhibits a dominant peak around 27°, attributable to the graphitic carbon (Supplementary Fig. 4) with no metallic Ni peak19,20, possibly due to the low content of Ni NPs while the Ni SACs are undetectable by XRD. The HRTEM examination of the carbon nanosheets free from Ni NPs reveals a mesoporous structure with a lattice distance of 3.4 Å, corresponding to the (002) facet of distorted graphitic carbon (Fig. 2d). The energy-dispersive X-ray spectroscopy (EDX) elemental mapping images of Ni@NiNCM show homogeneously distributed C, N, and Ni elements (Fig. 2e). Raman spectrum in Supplementary Fig. 5 displays three character peaks of D band (1,349 cm− 1), G band (1,583 cm− 1), and 2D band (2,650 cm− 1)21,22 with a ID/IG value of 1.01 for Ni@NiNCM, suggesting the existence of defects and/or disordered graphitic carbon structures associated with the N dopants and/or Ni NPs18.
X-ray photoelectron spectroscopy (XPS) analysis reveals the successful introduction of Ni, N, and C elements in Ni@NiNCM (Supplementary Fig. 6). Deconvoluted high resolution XPS N 1 s spectrum of Ni@NiNCM displays five typical peaks located at 398.2, 398.8, 399.7, 400.7, and 402.3 eV (Fig. 2f), attributable to pyridinic N, Ni-bonded N, pyrrolic N, graphitic N, and oxidized N, respectively. The high resolution XPS Ni 2p spectrum for Ni@NiNCM reveals the presence of metallic Ni0 (852.7 eV), Ni2+ (854.8 eV), and partially oxidized Ni3+ (855.8 eV)19,20 (Supplementary Fig. 7). The content of Ni species was determined to be 2.9 wt.% by inductively coupled plasma (ICP) analysis. Brunauer-Emmett-Teller (BET) measurements show a high BET surface area of 265 m2 g− 1 for the porous structure of Ni@NiNCM, attractive for an efficient exposure of the catalytic active sites and a favorable transfer of the ECR-relevant species (Fig. 2g)21,22.
ECR performance of Ni@NiNCM. The ECR activity of the Ni@NiNCM catalyst was evaluated in CO2-saturated 0.5 M KHCO3 electrolyte with three electrodes in an H-type cell. For comparison, ECR activity measurements were also performed on controlled samples of NiNCM prepared with sufficient acid leaching to remove Ni NPs from of Ni@NiNCM, Ni NPs, and Ni NPs/NiNCM (physical mixture). Details of the preparation procedures (see details in Supporting Information) and structure characterization for these controlled samples are present in Supplementary Figs. 8–12. As shown in Fig. 3a, Ni@NiNCM displayed a marked increase in current density in the CO2-saturated KHCO3 electrolyte compared with the Ar-saturated KHCO3, demonstrating the occurrence of ECR reaction under the CO2-saturated condition. The onset potential of Ni@NiNCM for ECR was determined to be -0.35 V from CO generation revealed by the gas chromatography (GC) profile in Supplementary Fig. 13 while no any liquid product was detected by 1H nuclear magnetic resonance (NMR) spectroscopy (Supplementary Fig. 14).
In order to evaluate the contribution of ECR catalysis to overall current density, the partial current densities of CO (jCO) for the as-prepared catalysts were measured. As can be seen in Supplementary Fig. 15, the Ni NPs displayed a negligible activity for CO2 reduction with jCO of about 0.01 mA cm− 2. In contrast, NiNCM delivered a more favorable ECR activity with the maximum jCO of 3.5 mA cm− 2 at a potential of -1.0 V. For Ni@NiNCM, the jCO dramatically increased up to 13.5 mA cm− 2 at -1.0 V, indicating that Ni NPs in Ni@NiNCM significantly enhanced the ECR activity of NiNCM, though pure Ni NPs or Ni NPs physically mixed with NiNCM (i.e., Ni NPs/NiNCM with jCO of 0.98 mA cm− 2) displayed almost no activity for CO2 reduction. Clearly, therefore, there is a strong synergistic effect between Ni NPs and NiNCM in the Ni@NiNCM catalyst. Among all the samples tested, Ni@NiNCM still exhibited the highest specific activity over the whole reduction potential even when the jCO was normalized to per electrochemical double-layer capacitance (Cdl) (Fig. 3b and Supplementary Fig. 16). Moreover, CO FE of Ni@NiNCM could maintain over 90% under the potentials ranging from − 0.7 V to -1.1 V, with the maximum CO FE of 97.6% at -0.9 V, while the CO FEs of NiNCM, Ni NPs, and Ni NPs/NiNCM were only 73.0%, 17.0%, and 21.1%, respectively, at -0.9 V (Fig. 3c). Notably, the observed high ECR performance of Ni@NiNCM exceeded that of almost all previously-reported benchmarking ECR electrocatalysts based on N-doped carbon, Ni SAC, and carbon-supported Ni NPs reported to date (Fig. 3d and Supplementary Table 1). The superb ECR performance of Ni@NiNCM was further supported by its turnover frequency (TOF) up to 839 h− 1 at -0.9 V (Supplementary Fig. 17 and Supplementary Table 2), a value which is almost two times higher than that of NiNCM (465 h− 1) and also superior to that of Ni NPs (25 h− 1).
To gain the insight of reaction kinetics, we obtained a Tafel slope of 90 mV dec− 1 for NiNCM, which is between 120 mV dec− 1 (Eqs. 1) and 60 mV dec− 1 (Eq. 2) and indicate that the *CO2− generation and the subsequent protonation reaction have a crucial effect on ECR process23. Compared with NiNCM, Ni@NiNCM showed a markedly decreased Tafel slope (52 mV dec− 1, Fig. 3e), indicating that the protonation step was significantly improved by Ni NPs in the Ni@NiNCM due to an enhanced proton capture. This is further supported by electrochemical impedance spectra (EIS) shown in Supplementary Fig. 18, in which Ni@NiNCM showed a smaller impedance, and hence the faster electron transfer than that of NiNCM and Ni NPs.
To evaluate the long-term stability, we tested simultaneously both the electrochemical stability and performance stability of Ni@NiNCM. As shown in Fig. 3f, the CO FE could maintain over 90% with the current density declined only by ~ 9% after 10-hours electrocatalysis. After the long-term testing, the morphology and chemical structure of Ni@NiNCM were found to be well-preserved from the post-test sample characterization by various microscopic and spectroscopic techniques, including TEM, XRD, and XPS. Clearly, therefore, Ni@NiNCM is highly stable under the ECR conditions (Supplementary Figs. 19–21).
Identification of active sites. In order to reveal the local chemical coordination environments of Ni@NiNCM, we further performed XAS measurements. The Ni K-edge X-ray adsorption fine structure (XANES) spectra show that the pre-edge of Ni@NiNCM is between NiO and Ni foil (Fig. 4a), indicating that the average valance of Ni species in the Ni@NiNCM is at a partially oxidized state. This is consistent well with the height of the white-line peak marked by dark red circle. Considering that the valance of the Ni NPs is zero, the higher oxidation state implies the existence of another coordination form of Ni atoms except the Ni NPs in Ni@NiNCM. The Fourier transforms of extended X-ray absorption fine structure (EXAFS) (R-space, Fig. 4b) for Ni@NiNCM displays a dominant peak similar to that of the Ni foil, corresponding to the Ni-Ni scattering path. The frontal position of Ni@NiNCM was offset to 2.1 Å, while that of the Ni foil was located at 2.2 Å, indicating that the Ni-Ni bond in the Ni@NiNCM was slightly shorter than that of the Ni foil. Interestingly, a slight acromion was observed for Ni@NiNCM referring to the Ni-N bonding of nickel phthalocyanine (NiPc), as marked by the light green shadow. Although the characteristic peak of Ni-N bonds was suppressed by the dominant peak of Ni-Ni in Ni NPs, the existence of Ni-Nx bonds in Ni@NiNCM could still be evident24, which was further supported by EXAFS fitting results in Supplementary Fig. S22 and Table 3. To obtain more precise information on the coordination environments of Ni species throughout the whole Ni@NiNCM catalyst sample, we applied the continuous wavelet transform (WT) to the EXAFS spectra (Fig. 4c). Compared with that of the Ni foil, the WT contour plot of Ni@NiNCM significantly shifted to the negative side due to the Ni-Nx contributions, further verifying the coordination of Ni and N. Impressively, the coexistence of the Ni-Nx coordination with Ni NPs attached on the carbon matrix could be directly observed by AC-STEM with EDX element mapping (Fig. 4d). The signals of Ni element in the EDX element mapping images could not only be clearly seen over the Ni NPs but also to be dispersed uniformly together with N element in the carbon matrix25, demonstrating that the Ni-Nx species were anchored around the Ni NPs. Furthermore, these atomic-level Ni sites around the Ni NPs could be clearly identified by atomic-resolution high-angle annular dark-field (HAADF) analysis, as shown in Figs. 4e and 4 f. The bright dots marked by yellow circles showed a distinctly different contrast with a diameter of ~ 0.22 nm corresponding to isolated Ni atoms. Thus, the above results clearly revealed that the atomic-level Ni-Nx sites anchored around Ni NPs on the carbon matrix.
To further verify the active sites, we used SCN− ions, as a remarkable catalytic activity inhibitor26, to block the atomic-level dispersed Ni-Nx sites in Ni@NiNCM. It was observed that the maximum CO FEs of Ni@NiNCM and NiNCM reduced from 97.6–71.9% (Fig. 4g and Supplementary Fig. 23) and from 73.0–51.9% (Supplementary Fig. 24), respectively, but not for the bare Ni NPs (Supplementary Fig. 25), upon the introduction of SCN− ions. These results revealed that the Ni-Nx sites played a significant role for the CO production during the ECR process, while the Ni NPs only accelerated reaction kinetics by boosting the protonation procedure (Eqs. 3 and 4, see above).
To obtain insights into the role of Ni NPs in accelerating protonation during the ECR process, we have conducted studies on the kinetic isotope effect (KIE) of H/D (H2O/D2O) over the Ni@NiNCM and NiNCM catalysts. The KIE can be used as an indicator of the proton transfer rate of water dissociation, which is the major proton donor for the ECR protonation process27. The KIE of NiNCM was calculated to be 2.45, while the KIE value over the Ni@NiNCM sharply decreased to 1.24. This result reveals that the introduced Ni NPs in the Ni@NiNCM is responsible for the dissociation of water to accelerate the proton transfer process (Supplementary Fig. 26a).
To gain further evidence for the role of Ni NPs, we have investigated the effect of local pH change that may contribute to the enhanced ECR performance. As such, three different electrolytes were used to adjust the local pH environment at the cathode/electrolyte interface with the following sequence: K2HPO4 < K2CO3 < K2SO428. The electrochemical results showed that the formation rate of CO product over both Ni@NiNCM and NiNCM catalysts increased in the same order of K2HPO4 < K2CO3 < K2SO4 (Supplementary Fig. 26b). However, the increased ratio of CO production rate for Ni@NiNCM and NiNCM (donated as RateNi@NiNCM/RateNiNCM ratio) in K2SO4 (2.66) is higher than those of in K2HPO4 (2.10) and K2CO3 (2.55) (Supplementary Fig. 26c). This observation demonstrates that the presence of Ni NPs accounts for the promoted CO production by accelerating the water dissociation process, although the dissociation of water becomes more difficult at a higher pH29,30. Furthermore, it was observed that the enhancement in the CO promotion rate over NiNCM was very limited with respect to that of Ni@NiNCM, further confirming the positive role of Ni NPs enhancing the dissociation of water to boost the proton transfer process.
DFT calculations. For clarifying the intrinsic ECR mechanism, we have constructed four different models based on the XAS results. These models are consisting of the NiN1, NiN2, NiN3, and NiN4 coordinated structures anchored adjacent to Ni NPs on the carbon matrix, which are donated as Ni@NiN1CM, Ni@NiN2CM, Ni@NiN3CM, and Ni@NiN4CM, respectively. As shown in Figs. 5a-5c, the detailed reaction pathways of CO2 transition to CO, hydrogen evolution traction (HER), and water dissociation in a weak alkaline solution (pH = 7.2) were calculated for those models. In Fig. 5a, the pathways for CO2 adsorption and CO desorption are almost up-hilled, indicating the needs of external potentials to drive the endothermic reactions. The free energies of CO2 adsorption for Ni@NiN1CM, Ni@NiN2CM, and Ni@NiN3CM models are 2.409, 2.713, and 3.389 eV (U = 0 V), respectively, while it is about 0.000 eV for Ni@NiN4CM. The NiN4 structure has been previously demonstrated to capture CO2 molecules31, while the Ni NPs could synergistically work for HER process32 which is in favor of adsorbing the Had and facilitates the protonation process of ECR. As such, the Ni NPs deposited on the carbon matrix can modify the electronic structures of adjacent NiN4 centers to form an interactive active center for CO2 reduction (Fig. 5a). For desorption of CO, the free energies are 2.866, 0.598, 2.290, and 2.567 eV for Ni@NiN1CM, Ni@NiN2CM, Ni@NiN3CM, and Ni@NiN4CM, respectively. The above four models display strong bonding forces of *CO intermediate on the active centers, indicating that the *COOH transition to *CO can be accelerated by the short bond length. The overpotentials of ECR process were calculated to be 2.866, 3.389, 2.713, and 2.567 eV for Ni@NiN1CM, Ni@NiN2CM, Ni@NiN3CM, and Ni@NiN4CM, respectively. These DFT results combined with the above PDOS calculations (Figs. 1d and 1e) reveal that the excellent electron transport and capture capacity of Had for the Ni@NiN4CM model leads to a significant decrease in the potential barrier for CO2 activation.
Figures 5b and 5c show the reaction pathway profiles for HER and water dissociation processes associated with these four models. As can be seen in Fig. 5b, the HER overpotentials are 1.755, 2.311, 2.151, and 0.746 eV for Ni@NiN1CM, Ni@NiN2CM, Ni@NiN3CM, and Ni@NiN4CM, respectively. Figure 5c shows the lowest free energy of 0.048 eV for Ni@NiN4CM during the water dissociation process. The weakened free energy for water dissociation can help to accelerate the H2O transition to Had, thus effectively capturing protons to boost the ECR reaction kinetics. In addition, the free energies of NiN4 model with *CO2, *COOH, and *CO intermediates are presented in Supplementary Fig. 27. These results indicate that Ni NPs can facilitate protons transport from the Ni NPs to NiN4 centers to form the intermediates, thus speeding up the whole ECR process. From Fig. 5d, it can be seen that the Ni NP can assist to capture H2O and Had, thus promoting the *CO2 transition to *COOH on the Ni atom in the NiN4 structure, and the proposed reaction mechanism is shown in Fig. 5e. Consequently, the surface state hybridization induced by the Ni NPs on NiN4 cause a change in the adsorption strengths of intermediates and potential barriers for certain elementary steps of ECR, which optimizes the reaction kinetics and boosts the ECR catalytic activities.
Zn-CO 2 battery. To demonstrate the ECR catalysts for potential applications, we developed an integrated Zn-CO2 battery based on the Ni@NiNCM cathode. A Zn-CO2 battery can safely convert CO2 into green chemical fuels and clean electricity. Figure 6a shows a schematic diagram of the Zn-CO2 battery developed in this study, in which the Ni@NiNCM was used as a cathode for efficient reducing CO2 to CO (also, NiNCM or Ni NPs/NiNCM was used for comparison) with Zn plate as the anode. Figure 6b shows the charge-discharge voltage profiles of the Zn-CO2 battery, along with the polarization curves for the anodic oxygen evolution of the Ni@NiNCM, NiNCM, and Ni NPs/NiNCM electrodes in Supplementary Fig. 2833–35. The power density of Ni@NiNCM was further calculated in Fig. 6c with a maximum value up to 1.0 mW cm− 2, which was much higher than that of NiNCM and Ni NPs/NiNCM. For discharge processes at different constant current densities (Fig. 6d), Ni@NiNCM displayed the higher discharge capacity, delivering a stable current of 1.0 mA cm− 2 at 0.22 V, along with an energy density of 386 Wh kg− 1. Figure 6e shows the galvanostatic discharge-charge cycling curve for the Zn-CO2 based on the Ni@NiNCM cathode at 1.0 mA cm− 2 for 90 circles (32 h), indicating a good durability. In addition, three Ni@NiNCM based Zn-CO2 batteries were put into series to light up a LED bulb (Fig. 6f). Furthermore, we have also successfully demonstrated the feasibility to use an integrated solar panel to recharge the Zn-CO2 battery (Figs. 6g and 6 h).