Introduction

Electrocatalytic reduction of carbon dioxide is a promising method to achieve carbon neutrality and mitigate the global warming crisis1,2. Transforming CO2-to-CO is a cost-effective method for electrocatalytic CO2 reduction, due to its practical manufacturing advantages in techno-economic assessments3. Researchers have noticed that several metals are capable of converting CO2 to CO, including gold (Au), silver (Ag), iron (Fe), cobalt (Co), and nickel (Ni), etc4,5,6,7. Among these catalyst candidates, Ni catalysts promise to provide high activity and FEco while suppressing the side reaction (hydrogen evolution reaction, HER), rendering Ni one of the most valuable and effective candidates for CO2-to-CO conversion8. Meanwhile, for practical application, the work potential range for CO2-to-CO conversion should be wide enough to match the high-fluctuating renewable electricity and to couple with different anode reactions. Therefore, the promising Ni catalysts should be capable of working in a wide range of potentials with highly intrinsic activity, FEco and stability9,10,11. Indeed, certain synthetic strategies have been reported to fabricate various Ni catalysts with different species/dispersions, morphologies and coordination environments12,13,14,15,16,17,18,19,20. For instance, Zhang reported that dual Ni atoms could offer a moderate Turnover frequency (TOF) of 77,500 h−1 per site with a potential window from −0.4 to −0.95 V (98–99% FECO)15. Also, Song reported that Ni-N species could achieve a TOF of 274,000 h−1 per site within a narrower potential range from −0.3 to −0.7 V (98–99% FECO)20. However, to satisfy the practical conditions, the Ni catalysts with the employment of current synthetic methods still can not offer sufficiently high intrinsic activity and FEco within a wide potential window. Therefore, it is urgent to develop a new synthetic strategy for fabricating a highly active, stable and selective Ni site for CO2-to-CO conversion, capable of operating within a sufficiently wide range of work potential.

In addition to the species/dispersions, morphologies, and coordination, the micro-environment formed by the vacancy around the active site could also play a key role in a synergistic effect, and even vacancy itself can be an active site in many catalytic processes21,22,23,24,25,26,27. For CO2-to-CO electrocatalytic conversion, there is barely reported work on simultaneous engineering of the vacancy and construction of highly pure and active sites for boosted CO2 conversion. Here we report a general strategy to synthesize the efficient Ni catalytic sites with uniform-large (UL) vacancies via the introduction of pre-designed atomically precise Ni clusters with sulfur ligands. Rather than the single role of atomically precise metal clusters reported previously as purely unsupported active sites with ligand-protecting, or precursors for the synthesis of supported ligand-protecting clusters13,28,29,30,31,32,33, in our method, they could serve entirely new dual roles in synthetic process, acting as a Ni precursor for Ni individual atoms and also etching the N (nitrogen)-rich carbon support like an in-situ nano “bomb” to form the uniform-large (UL) vacancy surrounding Ni atoms. With this strategy, highly pure and unsaturated NiNx species (“x” refers to coordination numbers) with UL vacancy could probably be fabricated via Ni cluster decomposition and in-situ etching of carbon support at high temperatures. As an expectation, this highly unsaturated NiNx species with UL vacancies could make these catalysts potentially more flexible to the adsorption changes of intermediates and promote this specific catalyst more stable, active and competitive to side reaction (HER). More strikingly, we found that this synthetic strategy demonstrated strong generality, where a series of atomically precise Ni clusters with different Ni atom numbers applied to the fabrication of this highly pure NiNx site with UL vacancy. As a consequence, with the employment of pre-deposition of atomically precise Ni clusters and in-situ sulfur-ligand-etching, we could fabricate a batch of highly pure, active, and stable NiNx sites with UL vacancies. Compared to the Ni-N species made by previously reported methods, this highly pure NiNx site with UL vacancy could dominantly suppress the side reaction (HER) and achieve almost 100% FEco and a high TOF of 350,000 h−1 per site within a wide potential range of about 1500 mV (from −0.12 to −1.6 V).

Results

Synthesis of atomically precise Ni clusters and NiNx catalysts

We first synthesized the atomically precise Ni clusters with a typical method34,35,36. Single-crystal X-ray diffraction over these different atomically precise Ni clusters ambiguously demonstrated that all these Ni clusters were ring-like species (inset images in Fig. 1a–f), which could facilitate the pre-deposition of the clusters on the carbon support. The sulfur ligands (including thiol/thiophenol) could behave as a precise in-situ etching agent to form a UL vacancy for the simultaneous Ni deposition. To extend the library of cluster precursors, we fabricated small Ni clusters (Ni5(SC2H4Ph)10 and Ni6(SC2H4Ph)12) and relatively larger clusters (Nin(SPhCH3)2n, n = 9–12) (Fig. 1a–f). Of note, for simplification, we used Ni5, Ni6, Ni9, Ni10, Ni11, and Ni12 as short forms for their molecular compositions. The as-obtained fresh clusters were determined by the high-resolution electrospray ionization mass spectrometry (ESI-MS) and Ultraviolet-Visible-Near Infrared (UV-Vis-NIR) spectroscopy (Fig. 1a–f and Supplementary Fig. S1). For instance, the dominant peaks at m/z ~ 1666.10, 1999.12, 2745.90, 3050.89, 3355.85, and 1831.93, respectively, could be assigned to the molecular ion peaks of [Ni5 + H]+, [Ni6 + H]+, [Ni9 + H]+, [Ni10 + H]+, [Ni11 + H]+, [Ni12 + 2H]2+ adducts (Fig. 1a–f), as evidenced by the well-matched high-resolution experimental MS spectra compared with the calculated isotopic MS patterns of each cluster (Fig. 1a–f). Subsequently, we selected small clusters (Ni5, Ni6) and relatively larger cluster (Ni10) as the deposition precursors to synthesize the NiNx sites with UL vacancies. The Ni cluster precursors were first grafted onto the nitrogen-rich carbon (NC) support. Then, the catalysts were further purified with ethanol to remove the physisorbed Ni clusters. Next, the as-prepared fresh Ni catalysts (designated as Nin/NC, “n” refers to the number of the atomically precise Ni clusters) were treated under high temperature to form UL vacancies by in situ sulfur-ligand-etching, providing a highly stable anchoring environment for the deposition of Ni atoms (Fig. 1g and Supplementary Table 1). We found that highly pure NiNx sites with UL vacancies were obtained through the deposition and pyrolysis of atomically precise Ni clusters along with in situ sulfur-ligand-etching.

Fig. 1: High resolution ESI-MS of the Ni clusters and the schematic illustration of NiNx fabrication from Nin clusters.
figure 1

a Ni5(SC2H4Ph)10, b Ni6(SC2H4Ph)12, c Ni9(SPhCH3)18, d Ni10(SPhCH3)20, e Ni11(SPhCH3)22, f Ni12(SPhCH3)24. In (af): the measured (black) and simulated (blue) isotopic distribution patterns of the corresponding molecular ion peaks. Inset (af): the molecular structures of the corresponding Ni clusters from Single-crystal X-ray Diffraction. g The schematic illustration of NiNx fabrication from Nin clusters. Note: The “x” in NiNx is about 2 revealed by the following characterization. Source data for (af) are provided as a Source Data file.

Characterization for Ni sites and UL vacancies

To shed light on the specific NiNx active sites with UL vacancies, we resorted to the atomic resolution aberration-corrected high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) imaging and scanning transmission electron microscopy electron energy loss spectroscopy (STEM-EELS) mapping to identify the local coordination environment around Ni atoms. For the fresh catalysts, Nin clusters were highly dispersed on NC support (Supplementary Figs. 24). The average particle sizes of Ni5/NC, Ni6/NC, and Ni10/NC were about 0.8, 1.0, and 1.3 nm (Supplementary Fig. 5). After pyrolysis at high temperatures, the Ni clusters were decomposed into individual atoms (Fig. 2a, b and Supplementary Figs. 610) and deposited on the vacancies formed by precise in situ sulfur-ligand-etching. No any visible clusters or nanoparticles were observed in the NiN2-Nin samples, demonstrating atomic dispersion of Ni species on NC support (Supplementary Figs. 610). Based on the quantitative atomic STEM-EELS elemental analysis (Fig. 2a, c, k), in the case of NiNx site derived from the small cluster (Ni6), it demonstrated that there were about two N atoms spatially correlated with one Ni atom, which confirmed Ni1-N2 coordination configuration. In other randomly selected areas, the atomic STEM-EELS mapping also demonstrated the identical coordination environment around the Ni atoms (Supplementary Fig. 11), suggesting the high purity of NiN2 sites derived from small Ni clusters (Ni6). For other small cluster (Ni5) derived NiN2-Ni5 sample, the coordination environment is the same as that of NiN2-Ni6 (Supplementary Fig. 12).

Fig. 2: Microscopic characterizations of NiN2-Nin (n = 6, 10) catalysts.
figure 2

HAADF-STEM images of the NiN2-Ni6 (a) and NiN2-Ni10 (b) samples. Atomic STEM-EELS mapping of Ni-L2,3 (green) and N-K (purple) in NiN2-Ni6 (c), NiN2-Ni10 (d) in corresponding select areas (marked in the yellow square in a, b). STEM-EELS mapping (e, g) and intensity profiles (f, h) of C-K for NiN2-Ni6 (c) and NiN2-Ni10 (d) respectively. Atomic STEM-EELS signals of N-K edge in i NiN2-Ni6 and j NiN2-Ni10. Quantitative STEM-EELS analysis of Ni/N molar ration based on Ni-L2,3 and N-K edge for k NiN2-Ni6 and l NiN2-Ni10, the Ni/N molar ratio is about 2 in NiN2-Nin catalyst. Source data for (il) are provided as a Source Data file.

Interestingly, the large Ni clusters (Ni10) demonstrated similar features with small clusters. The fresh Ni10 clusters were also highly dispersed on the NC, and the Ni atoms formed from Ni10 decomposition were also individually dispersed without any aggregation (Fig. 2b). As revealed by the quantitative atomic STEM-EELS elemental analysis, one Ni atom was possibly coordinated with two N atoms (Fig. 2d, l and Supplementary Fig. 13), which manifested a similar coordination structure with the catalysts derived from the larger clusters (Ni10) with an adequate homogeneity. It is worth noting that there were no sulfur atoms doped around the Ni atoms (Supplementary Fig. 14). For the coordinated nitrogen in the NiN2 site derived from both the small clusters (Ni6) and large cluster (Ni10), the STEM-EELS results (Fig. 2i, j) revealed that the nitrogen atoms were in sp2 hybridized form37,38, suggesting the same coordination environment around these NiN2 sites synthesized from different atomically precise Ni clusters. Interestingly, the fresh samples were oxidized by ozone at low temperature (denoted Nin/NC-O3), obtaining the mixtures of clusters and isolated atoms (Supplementary Figs. 1517). The NiN2 sites derived from Nin (Ni5, Ni6, and Ni10) were denoted as NiN2-Nin (n = 5, 6, 10). Also, NC with one cycle of Ni ALD (atomic layer deposition), pyrolysis with Ni(NO3)2·6H2O precursor and grafted with NiPc (Nickel Phthalocyanine) were denoted as 1cNi/NC, Ni/NC-T and NiPc/NC respectively for benchmark screening. For the benchmark Ni catalysts, Ni atoms were also atomically dispersed on the NC support (Supplementary Figs. 18, 19). In order to confirm the UL vacancy by sulfur-ligand-etching along with the Ni anchoring, we characterized the samples for more local information. As shown in the Fig. 2e–h and Supplementary Figs. 11e–f and 13e–f, the atomic STEM-EELS carbon mapping indicated that the local vacancy was about 0.6–0.7 nm per Ni atom, which was obviously larger and more uniform than that in normal electrocatalysts22,23,24,25,26,39,40,41. The existence of this UL vacancy aligned with the in situ sulfur-ligand-etching. To further confirm the process of in situ sulfur-ligand etching, we identified the formed gaseous species during this process by mass spectra. We observed some fragments of small sulfide molecules formed during this etching process (Supplementary Figs. 2022), suggesting that the etching of the NC support indeed occurred in this synthetic process.

In addition to the local information, we also resorted to more general characterizations. Similarly, the surface areas of the NC-related materials were totally different (Fig. 3a, b). The surface area of the NC was 888 m2/g, and the high-temperature annealed NC with small vacancies (designated NC-T) also offered a comparable surface area (712 m2/g). Interestingly, the NiN2-Nin samples derived from Ni5, Ni6, and Ni10 atomically precise clusters, demonstrated a doubled surface area (1490–1775 m2/g). The pore size distribution also suggested that 0.7–2.5 nm micropores increased sharply in these NiN2-Nin catalysts (Fig. 3b). The evolution of the pore size distribution confirmed that the UL vacancies were possibly formed along with the Ni deposition on the NC. It is worth noting that the formed pores around 0.7–2.5 nm were unambiguously consistent with the vacancy size revealed by atomic STEM-EELS mapping (Fig. 2e–h and Supplementary Figs. 11 and 13), which suggests the excellent uniformity of the vacancies around the Ni atoms in these NiN2-Nin catalysts. To further confirm the size of the vacancy formed in the synthetic process, we performed Small-angle X-ray Scattering (SAXS). As shown in the Fig. 3c and Supplementary Fig. 23, the pores from 0.7 to 2.5 nm also increased significantly, which aligns with the pore size evolution of BET and atomic STEM-EELS mapping. According to the X-ray diffraction (XRD) over these samples (Supplementary Fig. 24a), compared to the control materials, it should be noted that the NC support lost its certain crystalline morphologies along with the increasing disorder domains as revealed by the Raman spectroscopy (Supplementary Fig. 24b), which was probably ascribed to the in situ sulfur-ligand-etching. These results are also consistent with the atomic STEM-EELS mapping, surface area and trend of the pore size distribution. To further confirm the coordination environment of the as-prepared samples, we performed the X-ray absorption over these NiN2-Nin catalysts and other control samples (Fig. 3d, e) for more local coordination information. From the X-ray absorption near edge structure (XANES), the standard NiPc gave the highest valence state (+2) and the Ni foil showed a classic metallic state (Fig. 3d and Supplementary Fig. 25). The XANES spectra of all as-prepared catalysts located between the NiPc and Ni foil, indicating all catalysts derived from atomically precise Ni clusters were between 0 to +2 valence state. For fresh Ni5/NC, Ni6/NC, and Ni10/NC, they exhibited similar absorption features, suggesting that they shared the same local Ni-S coordination in these fresh samples. From the X-ray photoelectron spectroscopy (XPS) results, the fresh Nin/NC gave lower binding energy of the Ni 2p3/2 peak than NiN2-Nin (Supplementary Figs. 2628), which shows good agreement with XANES (Fig. 3d). Also, the binding energy of the Ni 2p3/2 peak for Nin/NC-O3 manifested high oxidation states (Supplementary Fig. 29).

Fig. 3: Characterizations of Ni-based catalysts.
figure 3

Nitrogen physisorption isotherm (a), and pore size distribution of NC, NC-T, and NiN2-Nin (n = 5, 6, 10) samples (b). c Small-angle X-ray Scattering (SAXS) of NC, NC-T, and NiN2-Nin (n = 5, 6, 10) samples. X-ray absorption of various Ni catalysts and standard materials, XANES (d) and EXAFS (e) spectra and fitting data for all related Ni materials. Source data for Fig. 3 are provided as a Source Data file.

After high-temperature treatment for these fresh catalysts, the obtained NiN2-Ni5, NiN2-Ni6, and NiN2-Ni10 catalysts exhibited almost identical characteristic features and similar curves, suggesting the similar Ni-N species in these samples (Fig. 3d). In addition, the fitted Ni–N peak in the N1s XPS spectra (Supplementary Figs. 3033) also suggested the existence of Ni–N coordination in these NiN2-Nin samples. The extended X-ray absorption fine structure (EXAFS) was also investigated on all Ni catalysts for the local coordination environments (Fig. 3e). For the standard materials, the peaks located at 1.41 Å (NiPc) and 2.11 Å (Ni foil) were assigned to Ni-N and Ni-Ni coordination shells respectively. For fresh catalysts without high-temperature treatments, the experimental EXAFS curves of all Nin/NC catalysts almost demonstrated the same peak around 1.63–1.65 Å, which corresponded to the Ni-S coordination shell. The fitting curves (dashed line in Fig. 3e) with the Ni-S4 pathway showed excellent agreement with the experimental results. Interestingly, after the fresh catalysts were pretreated with ozone at low temperature (Nin/NC-O3), there were two main peaks between 1.41 and 1.61 Å, indicating there were Ni-S and Ni-O/N coordinations in these samples (Supplementary Figs. 34). After the high-temperature treatment, all NiN2-Nin catalysts gave identical curves with the same main peak around 1.41–1.43 Å, suggesting that all NiN2-Nin catalysts obtained the highly identical Ni-N coordination environment. Also, the fitting curves with Ni1-N2 pathway over these NiN2-Nin samples are highly consistent with the experimental curves (dashed line in Fig. 3e, Supplementary Fig. 35 and Table 2), and this NiN2 coordination from EXAFS also agrees well with the results of atomic STEM-EELS mapping. The Ni catalyst with a potentially small vacancy (1cNi/NC) also showed a main peak around 1.47 Å. The XAFS spectra of Ni/NC-T were also provided for benchmarking (Supplementary Figs. 34, 35).

Taken together, the NiN2 site with UL vacancy could be fabricated by the employment of Ni5, Ni6 and Ni10 atomically precise clusters, providing a general strategy to synthesize highly pure and active NiN2 sites with UL vacancies.

Evaluation of CO2-to-CO activity

Electrocatalytic CO2-to-CO tests over all the catalysts were first performed in the H-type cell (Fig. 4a), and the catalysts (Nin/NC) were also screened with treatments at different temperatures (Supplementary Figs. 3639). The only products were the CO and H2 in all cluster-related catalysts, no liquid products were detected as revealed by 1HNMR spectra after the tests (Supplementary Figs. 4041). As expected, the carbon itself just produced H2 as the main product. For the 1cNi/NC (small vacancy) and NiPc/NC (NiN4 site) catalysts, they just offered a 300 mV of potential range with 80% FEco, and ~200 mV of potential range with 60% FEco, respectively. For the catalysts derived from normal Ni precursor (Ni(NO3)2·6H2O) without in situ sulfur-ligand-etching (Ni/NC-T), it only provided a narrow potential range with low FEco. For NiN2-Nin catalysts derived from the Nin clusters (Ni5, Ni6, and Ni10) could display a 1000 mV of potential range with ~99% FEco, demonstrating a high CO selectivity and a wide potential range. All the tests in the flow-cell mode were also performed as shown in Fig. 4b–h and Supplementary Figs. 4250. The fresh Nin/NC (Ni5, Ni6, and Ni10) demonstrated a low FEco (less than 20%), indicating the Nin cluster itself was retarded to the CO2 reduction. Also, we tested the ozone pretreated catalysts, and they were highly active for the side reaction (HER) rather than for CO2-to-CO reaction (Supplementary Fig. 47). However, the NiN2-Ni6 with UL vacancy presented above 99% FEco within almost 1500 mV (from −0.12 to −1.6 V) of potential range (Fig. 4b and Supplementary Fig. 49a). Interestingly, it showed the same scenario in the NiN2-N5 and NiN2-N10, demonstrating about 99% FEco with about 1400 mV of potential range (from −0.2 to −1.6 V). Therefore, these NiN2-Nin catalysts derived from different atomically precise Ni clusters could possibly obtain the highly identical active site (NiN2), demonstrating similar electrocatalytic performance for CO2-to-CO process. Compared with the catalysts synthesized by other methods (Fig. 4c and Supplementary Table 3), NiN2-Nin fabricated through the current synthetic protocol demonstrated wide potential window (FEco > 95%). For the cathode energy efficiency (CEE), the fresh and ozone-treated catalysts both gave a low CCE (Supplementary Fig. 49c). In contrast, the NiN2-Nin catalysts could offer an ideal CEE value from −0.12 to −1.6 V, stepping close to the theoretical value of CEE in this system (Supplementary Fig. 49c). The NiN2-Nin could also obtain high (~100%) FEco at high work current (300–1200 mA), suggesting that the NiN2 sites with UL vacancies derived from different Ni clusters exhibited boosted inhibition towards HER (Fig. 4d, e and Supplementary Fig. 49b). For CO2 conversion, the NiN2-Nin could demonstrate 7–20 times higher than fresh and/or ozone-treated catalysts (Supplementary Fig. 50). 68% of CO2 conversion could be achieved with 1 mg NiN2-Nin catalysts at a current of 900 mA for all NiN2-Nin catalysts (Fig. 4f). Compared to other reported Ni catalysts (Fig. 4g), we found that the TOFs in the NiN2-Nin catalysts were definitely the highest, achieving about 350,000 h−1 in these NiN2-Nin samples. The stability tests were also carried out on the all NiN2-Nin catalysts (Fig. 4h and Supplementary Fig. 51), and the catalyst could maintain high FEco (>99%) for more than 50 h without any obvious deactivation or aggregation, and 116 h stability in 0.5 M KHCO3 at −0.6 V (Supplementary Figs. 52, 53), suggesting that the NiN2-Nin could not only give a high TOF, CO2 conversion and FEco across a wide potential range but also display excellent stability in the electrocatalytic CO2-to-CO process.

Fig. 4: Electrocatalytic CO2-to-CO test.
figure 4

a The FE values of CO for NiN2-Nin (n = 5, 6, 10) and various Ni-based reference catalysts in 0.5 M KHCO3 with H-type Cell, b The FE values of CO for NiN2-Nin in 1 M KOH in flow cell (1 cm2 GDE). c Comparison of work potential window with other reported catalysts for high FEco in flow cell. d Linear sweep voltammetry (LSV) curves, e FE values of CO and f CO2 conversions for 1 mg NiN2-Nin (1.5 mg for NiN2-Ni10) on 5 cm2 GDE in 1 M KOH in flow cell. g Comparison of apparent TOFs of CO generation with other reported Ni catalysts tested in flow cell in 1.0 M KOH. h Time-dependent FEs and potential of NiN2-Nin during constant current electrolysis at 100 mA. The points marked by arrows in the (h) represent the refreshing. Note: The details and related references for (c, g) are listed in the Supplementary Table S3. All experimental potentials were not iR corrected. Source data for Fig. 4 are provided as a Source Data file.

Catalytic mechanism over Ni2-Nin catalysts

In-situ attenuated total reflection Fourier transform infrared spectroscopy (ATR-FTIR) was performed on these NiN2-Nin samples to deeply investigate the reaction intermediates to uncover the electrocatalytic CO2-to-CO process occurring on the NiN2 site with UL vacancy. The pathway of electro-catalytic reduction of CO2-to-CO was reported in previous work, and the adsorbed intermediates are COOH* and CO*42,43. As reported in previous work44,45,46,47, the peak at 2100 cm−1 was assigned to CO* vibration on NiN2-Ni5, and peaks located at 1671 and 1428 cm−1 were ascribed to the adsorption of COOH* (Fig. 5a). Also, the putative adsorbed H2O* (1639 cm−1) and HCO3* (1364 cm−1) species were also observed in the NiN2-Ni5 catalyst. As expected, in the NiN2-Ni6 and NiN2-Ni10 catalysts, the adsorbed species were almost identical (Fig. 5b, c), suggesting that same active sites and electrocatalytic reduction mechanism were present in these NiN2-Nin catalysts. These in-situ ATR-FTIR results were consistent with the atomic STEM-EELS mapping and XAFS results. Furthermore, we also performed the quasi in-situ XPS (Supplementary Fig. 54) on NiN2-Nin catalysts to elucidate whether there are the electronic changes of Ni atoms during the CO2-to-CO conversion. The quasi in-situ XPS spectra demonstrated no obvious electronic changes of Ni atoms in NiN2-Nin catalysts, suggesting that the catalysts reconstruction may not occur in these systems under reaction condition.

Fig. 5: In situ measurements and theoretical calculations.
figure 5

ac In situ ATR-FTIR over NiN2-Nin (n = 5, 6, 10) at different potentials in 0.5 M KHCO3. d Schematic illustrations of reaction intermediates on optimized NiN2 with large vacancy (NiN2-Nin). Ni, N, O, C, and H atoms are in green, blue, red, gray, pink, respectively. e Calculated free-energy diagram for CO2 reduction to CO in different models. f Comparison of GH*-GCO2* at different applied potentials (U) between NiN2-Nin and NiN2-SV. Note: The value of GH*-GCO2* could reveal the competitive relationship between HER and CO2 reduction, especially for the adsorption of H and CO2 at the first step. Source data for (ac) and (e, f) are provided as a Source Data file.

To further elucidate the high performance and the mechanism of the NiN2-Nin catalysts with UL vacancies for electrocatalytic CO2-to-CO conversion, we performed density functional theory (DFT) calculations. To investigate the vacancy around the NiN2 site, we constructed three structures, including NiN2 site with a uniform-large vacancy (designated NiN2-Nin), NiN2 site with small vacancy (designated NiN2-SV), and four-coordinated Ni as benchmark (designated NiPc/NC), and the corresponding adsorption intermediates are also provided (Fig. 5d, Supplementary Figs. 5557). As shown in Fig. 5d, all adsorbed intermediate species on NiN2 with large vacancy are displayed. The process of CO2 reduction to CO is generally believed to involve different adsorbed intermediates, including CO2*, COOH* and CO*. We depicted the changes in Gibbs free energy by calculating the adsorption free energies of these intermediates as shown in Fig. 5e. For benchmark model (NiPc/NC in Fig. 5e), the rate-determining step (RDS) is the formation of the COOH* intermediate. Compared to four-coordinated Ni (NiPc/NC), the enhanced adsorption capability of Ni atoms shifted the RDS from the formation of the intermediate (COOH*) to the desorption of products (CO*) in NiN2-Nin and NiN2-SV. Based on the RDS in the CO2-to-CO process (Fig. 5e and Supplementary data 1), the catalytic performance of NiN2-Nin (RDS = 0.92 eV) was significantly better than that of NiPc/NC (RDS = 1.78 eV) and NiN2-SV (RDS = 1.32 eV). A larger 8×8 graphene supercell was also simulated for NiN2-Nin, which revealed that the free energy of each intermediate varies by ~0.2 eV or less (Supplementary Fig. 58). Furthermore, compared to NiN2-SV, the UL vacancy around the NiN2 site helped protect the intermediates from the influence of other atoms in the carbon framework of the catalyst except for the active Ni atoms, ultimately leading to a weakened CO* adsorption strength. Moreover, the “d orbital center” (Supplementary Fig. 59) of NiN2-Nin (−1.12 eV) located between that of NiPc/NC (−1.78 eV) and NiN2-SV (−0.82 eV), further suggesting that the adsorption capacity of NiN2-Nin was between that of NiPc/NC and NiN2-SV. This observation was consistent with the trend of COOH* and CO* adsorption. In the electrocatalytic CO2-to-CO process, the HER is the most competitive side reaction. As shown in Fig. 5f, we illustrated a competitive relationship between HER (Supplementary Figs. 60, 61) and CO2 reduction at different applied potentials (U). In this context, the ordinate was defined as GH-GCO2, where a more positive value indicated a greater tendency for CO2 adsorption. We observed that the value of GH-GCO2 in NiN2 site with large vacancy (NiN2-Nin) was significantly higher than that in NiN2-SV, demonstrating higher selectivity for CO2 reduction in the NiN2 system with large vacancy within a wide potential. Compared to small vacancy (NiN2-SV), this trend confirmed that UL vacancy could assist the NiN2 site in facilitating the reaction towards the CO2-to-CO pathway across the entire potential range. The changes of reaction energies in NiN2 with large vacancy (NiN2-Nin) with different bias were also provided in the Supplementary Fig. 62. Taken together, the theoretical results reveal that the large vacancy around NiN2 could boost the activity, FEco and the stability in wide potential range. This agrees well with the observed intermediates results from in-situ infrared experiments, and the experimental catalytic performance.

Discussion

In this work, with the employment of atomically precise Ni clusters, we developed a general synthetic strategy for the synthesis of highly pure NiN2 sites with UL vacancy via the “pre-deposition+ pyrolysis”, and in-situ precise sulfur-ligand-etching. The atomic STEM-EELS mapping, N2 physisorption, SAXS, and XAFS characterizations confirm the NiN2 species and UL vacancies around the Ni atoms. In the electrocatalytic CO2-to-CO conversion, this NiN2 with UL vacancy demonstrates highly electrocatalytic performance, achieving the highest TOF of ~350,000 h−1 and ~100% FEco within a wide potential range (~1500 mV). DFT and in-situ ATR-FTIR further elucidate that the UL vacancies around the NiN2 species could promote the durability at high potential with high CO selectivity, and facilitate the desorption of the target products (CO). This original synthetic strategy provides a promising method to engineer the vacancy with a highly active site by using atomically precise metal clusters towards highly catalytic performance in other reactions.

Methods

Chemicals and materials

Ni(NO3)2·6H2O (Aladdin, 99.99%), Zn(NO3)2·6H2O (Aladdin, 99.99%), NiCl2·6H2O (Aladdin, 99.9%), 4-methylphenthiophenol (Aladdin, 98%), tetrahydrofuran (Aladdin, 99.99%), toluene (Aladdin, 99.8%), C5H8N2 (Aladdin, 99.99%), Methanol (Sinopharm Chemical Reagent Co., Ltd, ≥99.8%), Ni(C5H5)2 (Aladdin, NiCp2, 98%), anion-exchange membrane (Fuelcellstore, FAA-PK-130), ethanol (Beijing Chemical Works, ≥99.8%), HCl (Sinopharm Chemical Reagent Co., Ltd, 36%-38%), sulfuric acid (H2SO4, Aldrich, 98%), Nafion solution (Sigma-Aldrich, 5 wt%), N, N-Dimethylformamide (Sinopharm Chemical Reagent Co., Ltd, ≥99.8%), 2-phenylethanethiol (Aladdin, 97%), NaBH4 (Sinopharm Chemical Reagent Co., Ltd, 96%), CH2Cl2 (Aladdin, HPLC). All gases in the experiments were purchased from Nanjing Special Gas Co., Ltd.

Synthesis of NC support

Solution A: 3.0 g Zn(NO3)2·6H2O dissolved in 50 mL CH3OH, solution B: 6.5 g dimethylimidazole dissolved in 100 mL CH3OH. Solution B was added dropwise to A and vigorously stirred for 24 h to obtain a white uniform suspension. ZiF-8 sample was acquired by centrifuge and rinsed three times with methanol and then dried at 60 °C in a vacuum. Then, 0.4 g ZiF-8 was pyrolyzed at 1100 °C for 6 h under the protection of Ar to obtain NC support.

Synthesis of NC-T

To acquire the blank control material, 40 mg NC support was ground evenly with 2.4 g melamine and then pyrolyzed at 950 °C for 1 h in Ar (99.999%, 30 ml/min) atmosphere.

Synthesis of Nin clusters

Synthesis of Nin (PET)2n (n = 5, 6; PET, phenylethylthiol, SC2H4Ph) clusters

100 mg of nickel(II) chloride hexahydrate (NiCl2·6H2O) was dissolved in a 9 ml mixture of tetrahydrofuran (THF) and methanol (MeOH) with a 2:1 ratio. Subsequently, 282 μL of PET was introduced into the solution and stirred for 10 min. 2 ml of an aqueous solution containing 40 mg of sodium borohydride (NaBH4), pre-chilled to 0 °C, was rapidly added to the above solution. After the gas evolution, the suspension was stirred at 1000 rpm for 3 h. Residual reactants were removed via twice washing with methanol, and the product was subsequently extracted using dichloromethane (DCM) and stored at −8 °C. Ni5(PET)10 and Ni6(PET)12 were then isolated and purified through preparative thin-layer chromatography (PTLC) employing a polar silica-gel stationary phase and a mobile phase consisting of mixture of petroleum ether and DCM (volume ratio ~ 2:1).

Synthesis of Nin(4MPT)2n (n = 9,10,11,12; 4MPT, 4-methylphenthiophenol, SPhCH3) clusters

24 mg of nickel(II) chloride hexahydrate (NiCl2·6H2O) were rapidly dissolved in 20 ml ethanol to form a verdant solution. After 15 min of stirring, 15 mg of 4MPT were introduced into the solution, and the mixture was stirred continuously for 30 min, followed by a fast addition of 3 ml of an ethanol solution containing 13.7 mg of sodium borohydride (NaBH4). The reaction proceeded for 12 h at ambient temperature. After the completion of the reaction, the crude products were harvested via centrifugation and triple ethanol washing. Subsequently, nickel cluster homologs were extracted from the crude cluster product with 2 ml CH2Cl2. Thin-layer chromatography (TLC) was employed for the separation of nickel cluster homologs, with a developing solvent comprising a mixture of CH2Cl2 and petroleum ether with a 1:3 volume ratio.

Synthesis of Nin (n = 5, 6, 10)/NC

5 mg Nin (n = 5, 6, 10) cluster were dissolved in 20 ml dichloromethane, then 100 mg NC support was added to the above solution. The mixture was stirred at room temperature (25 °C) for 12 h and then filtered after the addition of 200 ml ethanol to obtain Nin/NC precursor. The residual organic solvent was removed by vacuum drying at room temperature.

Synthesis of Nin (n = 5, 6, 10)/NC-O3

The synthesis of Nin (n = 5,6,10)/NC-O3 catalysts were performed in a viscous ALD flow reactor (ACME Beijing Technology Co., Ltd) at a pressure of 0.6 Torr. Typically, 100 mg Nin /NC (n = 5, 6,10) powders were placed in the ALD chamber at 150 °C. 50 ml/min of N2 carrier gas and 150 ml/min of ozone (O3) were dosed into the chamber. The timing sequence was 300 s and 600 s for O3 exposure and N2 purge, respectively.

Synthesis of NiN2-Nin (n = 5, 6, 10)

Ni6/NC was pyrolyzed at different temperatures, 40 mg of Ni6/NC and 2.4 g of melamine were ground evenly, and then pyrolyzed at 750 °C, 850 °C, 950 °C and 1050 °C for 1 h under the protection of Ar (99.999% 30 ml/min) to obtain NiN2-Ni6. To obtain NiN2-Ni5 and NiN2-Ni10, 40 mg of Ni5/NC and Ni10/NC were ground evenly with 2.4 g melamine and then pyrolyzed at 950 °C for 1 h in Ar (99.999% 30 ml/min) atmosphere.

Synthesis of 1cNi/NC

Ni ALD was performed in a viscous ALD flow reactor (ACME Beijing Technology Co., Ltd) at a pressure of 0.6 Torr. Ultrahigh-purity N2 (99.999%, Nanjing Special Gases) at a flow rate of 150 ml/min was used as the carrier gas. The timing sequence was 100 s, 180 s, 600 s, 300 s for NiCp2 exposure, N2 purge, O2 exposure, and N2 purge, respectively.

Note: 1cNi/NC, 3 N-coordinated Ni species with small vacancy, was also provided for the comparison with NiN2-Nin (Supplementary Fig. 63).

Synthesis of Ni/NC-T

5 mg Ni(NO3)2 and 100 mg NC support were dissolved in 20 ml H2O, then the mixture was stirred at room temperature for 24 h and then filtered to obtain Ni/NC precursor. 40 mg of Ni/NC and 2.4 g of melamine were ground sufficiently and then pyrolyzed at 950 °C for 1 h under Ar atmosphere to obtain Ni/NC-T.

Synthesis of NiPc/NC

5 mg NiPc were dissolved in 20 ml N, N-Dimethylformamide (DMF), then 100 mg NC were added to the above solution, followed by centrifuging and washing with DMF three times, finally dried at 80 °C to obtain NiPc/NC catalyst.

Characterization

Inductively Coupled Plasma Optical Emission Spectrometer (ICP-OES) analysis was performed on an iCAP 7400 to determine the elemental percentage. X-ray diffraction patterns were collected with a Rigaku TTRΙΙΙ diffractometer with Cu Kα radiation at room temperature. Ex situ X-ray photoelectron spectroscopy (XPS) was operated on a Kratos Axis supra+ spectrometer with an excitation source of monochromatized Al Ka (hν = 1486.6 eV) and a pass energy of 40 eV. Raman results were acquired on LabRAM HR Evolution Laser Confocal Raman Microscope. The XAFS spectroscopy results of Ni K-edge are collected at the X-ray absorption fine structure for the catalysis beamline of the Singapore Synchrotron Light Source, Singapore, and a nickel foil is used to calibrate the energy. AC-HAADF-STEM and EDS mapping were operated on a JEOL-F200 transmission electron microscope operating at 200 kV. Atomic STEM-EELS analysis was performed on the Nion HERMES-100 equipped with a C3/C5 corrector, operated at 60 kV with a convergence angle of 32 mrad. The collection angle is 0–75 for STEM-EELS and 75–210 mrad for HAADF. The probe current for STEM-EELS spectrum image acquisition is about 8–17 pA. Principal component analysis, together with Savitzky-Golay smooth was conducted as the denoising strategy to improve the signal-to-noise ratio for STEM-EELS mapping. It should be noted that all the signals/spetra related to the N hybridization state analysis and the quantitative STEM-EELS analysis were the raw data without processing. The quantification was achieved by Digital Micrograph through the following equation:

$$\frac{{N}_{{Ni}}}{{N}_{N}}=\frac{{I}_{{Ni}-L}}{{I}_{N-K}}\times \frac{{\sigma }_{N-K}}{{\sigma }_{{Ni}-L}}$$
(1)

where N is the areal concentration of the atoms, I is the integrated signal intensity, and σ is the cross-section for ionization of an electron in the associated shell. The Hartree-Slater model was chosen to calculate the theoretical cross-section. The region of 40 eV near the edge was excluded and the following region of 70 eV was taken into account as the integration region. (University of Chinese Academy of Sciences)

In situ attenuated total reflection Fourier transform infrared spectroscopy (ATR-FTIR) was conducted on a Perkin Elmer Spectrum 3 Spectrometer equipped with a mercury cadmium telluride (MCT) detector. The H-type electrochemical cell consisted of a working electrode with a catalyst loaded onto an ATR crystal (Au-coated Si semi-cylindrical prism), a high-purity graphite counter electrode (separated by Nafion 117 membrane), an Ag/AgCl reference electrode, and CO2-saturated 0.5 M KHCO3 solution as the electrolyte, which was installed in the spectrometer. Electrolysis was performed over a potential range from the open circuit potential (OCP) to −1.1 V vs. RHE, with spectra recorded at each potential after stabilization. N₂ physisorption analysis were conducted at −196 °C using a Micromeritics ASAP 2460 surface area and porosity analyzer to ascertain the specific surface area. Preceding the analysis, the samples were degassed at 200 °C for a minimum duration of 12 h. Small-angle X-ray scattering (SAXS) were performed on Xeuss 2.0 (Xenocs, France) with a Cu X-ray source (wave-length of 0.1542 nm) and a detector of Pilatus 300 K (Dectris). Heating experiments of 20 mg Ni6/NC powder were conducted under He flow (30 mL/min) using a thermogravimetric analyzer (Netzsch STA449F3) coupled to a mass spectrometer (Netzsch QMS 403Q).

Quasi in-situ XPS spectra of all electrodes completing CO2 electrolysis in CO2 saturated 0.5 M KHCO3 solution at different potential, were collected on a Thermo Scientific ESCALAB Xi+ X-ray photoelectron spectrometer (the Vacuum Interconnected Nanotech Workstation of the Suzhou Institute). The transfers of electrodes to XPS vacuum chamber were protected with the inert argon.

Electrochemical measurements

The electrochemical measurements were operated on the CorrTest-CS310X workstation (Wuhan Corrtest Instrument Corp. Ltd). During the H-type electrolytic cell test, the catalyst ink consists of 5 mg catalyst, 950 μl isopropanol, and 50 μl Nafion solution. 0.5 mg catalyst deposited on 2 cm2 glassy carbon sheet (0.25 mg/cm2) was employed as the working electrode, which was immersed in 0.5 M KHCO3 (1 cm2 platinum foil as counter electrode and Ag/AgCl as reference electrode). All the electrolytes were prepared and used immediately. For CO2 CO2-saturated 0.5 M KHCO3 solution, the value of pH was 7.20 ± 0.10. The Ag/AgCl reference electrode was calibrated in a three-electrode cell employing dual Pt foil electrodes (as working/counter electrodes) and H2-saturated electrolyte (1 h purging), with cyclic voltammetry conducted at 1 mV s−1. Nafion117 cation exchange membrane (size: 3 cm × 3 cm, thickness: 178 µm) separates the anode and cathode chambers. Nafion-117 membranes were pretreated by heating in Milli-Q water at 80 °C for 1 h, followed by three repetitions of sequential cycles (For each cycle: 3% H2O2 and 1.0 M H2SO4 treatments at 80 °C for 1 h with intermediate Milli-Q water rinsing). Finally, the membranes were stored in fresh Milli-Q water for further use.

For experiments in the flow cell, catalyst ink contains 5 mg catalyst, 3 ml ethanol, and 100 μl Nafion (sigma, 5%) solution. A piece of gas diffusion electrode (GDE, 1 × 1 cm2) was loaded with 0.2 mg catalysts, or a 5 cm2 GDE was loaded with 1 mg catalysts (1.5 mg for NiN2-Ni10) as the cathode (Ag/AgCl as reference electrode). A mass flow controller controlled gaseous CO2 entering the gas chamber and then flowing to the gas chromatograph at a flow rate of 10 ml/min monitored by a mass flowmeter. 1 M KOH at a constant rate of 10 ml/min was circulated through the cathode and anode chamber (nickel foam as counter electrode), which were separated by the FAA-3-PK-130 anion exchange membrane (size: 1 cm × 1 cm or 3 cm × 3 cm, thickness: 130 µm). The FAA-3-PK-130 anion exchange membrane was fully converted to OH⁻ form by soaking in 1.0 M KOH at room temperature for 24 h, then rinsed with ultrapure water to neutral pH (~7). The pretreated membrane was then stored in ultrapure water before further use. All experimental voltages were not iR corrected. For 1 M KOH solution, the value of pH was 13.9 ± 0.10. All potentials reported in this study were converted to the reversible hydrogen electrode (RHE) scale using the following equation:

$${{{\rm{E}}}}({{{\rm{vs}}}}.{{{\rm{RHE}}}})={{{\rm{E}}}}({{{\rm{vs}}}}.{{{\rm{Ag}}}}/{{{\rm{AgCl}}}})+{{{\rm{E}}}}^{{{\rm{\theta }}}}({{{\rm{Ag}}}}/{{{\rm{AgCl}}}})+0.0592\times {{{\rm{pH}}}}$$
(2)

Products analysis

During the electrocatalytic reduction process, the reduction products of CO2 were analyzed online by gas chromatography (GC9790 Plus, Zhejiang Fuli Analytical Instruments Co., Ltd) and the gaseous products entered into a 1 ml quantitative loop at a flow rate of 10 ml/min. Among them, H2 and high-concentration CO are quantified by a thermal conductivity detector, and low-concentration carbon monoxide is quantified by a flame ionization detector, the correlation coefficient of H2 and CO calibration curves is 99.99% (Supplementary Figs. 64, 65). Faradaic efficiency (FE) of H2 and CO was calculated by the following equation.

$${FE}=\frac{n\times F\times W\times v\times P}{R\times T\times I}$$
(3)

Where, n is the electron transferred during the reaction of carbon dioxide and water to the product (CO and H2, n = 2). F = 96485 C · mol−1; W is the volume fraction of product confirmed by calibration curves from standard gas determined by gas chromatography; v is the gas flow rate through the quantitative loop of gas chromatography (mL/min); P = 1.01 × 105 Pa; R = 8.314 J mol−1 K−1; T = 273.15 K. I refers to the total current of cell.

TOF was calculated by the following equation.

$${TOF}=\frac{{I}_{{CO}}\times {M}_{{Ni}}}{{n\times F\times \beta \times m}_{{cat}}}\times 3600$$
(4)

Where, ICO (A) is the partial current of CO; MNi is the atomic mass of Ni (58.69 g/mol); n is the electrons transferred in reaction; F = 96485 C · mol−1; β (wt%) is the amount of Ni loading in Ni based catalysts from ICP-AES analysis; mcat (g) is the mass of catalyst on electrode.

The cathode energy efficiency (CEE) of CO is calculated as follows:

$${CEE}=\frac{{FE}\times {E}_{0}}{1.23-V}\times 100\%$$
(5)

Where E0 is the thermodynamic cell potential of CO2 reduction to CO at the cathode (E0 = 1.23 − (−0.106) = 1.336 V); the thermodynamic potential of H2O oxidation at the anode is 1.23 V and equilibrium electrode potential of CO2 reduction to CO is −0.106 V. V is the applied potential of a full cell without iR correction.

Single pass CO2 to CO Conversion (SPCC) of different catalysts were calculated as follows:

$${SPCC}=\frac{{I}_{{CO}}\times 60(s)\times {1000\times V}_{{{{\rm{m}}}}}}{n\times F\times v\times 1(\min )}$$
(6)

ICO (A) is the partial current of CO; Vm (24.05 L/mol) is the molar volume of gas at 25 °C; n is the electrons transferred in reaction; F = 96485 C · mol−1; v (mL/min) is the gas flow rate through the quantitative loop of gas chromatography.

1H NMR spectra were collected on AVANCE 400 MH Superconducting Fourier Nuclear Magnetic Resonance Spectrometer. Liquid NMR samples consist of 400 µl of electrolyte collected after electrocatalysis, 100 µl of D2O as the solvent, and 100 μl DMSO as an internal standard.

Computation details

Density-functional theory (DFT) calculations were performed with the Vienna Ab-initio Simulation Package (VASP) codes 5.448. PAW pseudo-potentials49 with projectors up to l = 2 for Ni and l = 1 for C, N, O, and l = 0 for H, and the revised Perdew-Burke-Ernzerhof (RPBE) exchange-correlation functional50 were employed in our calculations. A plane wave cutoff of 500 eV was applied in the calculations.The valence electron numbers for Ni, C, N, O, and H are 10, 4, 5, 6, and 1, respectively. The convergences of energy and force were set to 10−4 eV, and 0.02 eV/Å, respectively. Spin-polarized calculations with Gaussian smearing in combination with a width of the smearing 0.05 eV were identified for all the analyzed structures. The spin polarization of Ni in NiN2-Nin were initialized at 0, 1, 2, 3, and 4 unpaired electrons, where all but 0 resulted in a net spin of 1 unpaired electron after optimization, which also yields the lowest energy. The initial magnetic moments of Ni in our calculations were set to one. The calculated structure was set by a 6 × 6 graphene supercell, and we transformed the graphene rhombic unit cell into a rectangular unit cell through coordinate transformation. A vacuum layer of 20 Å and a 3 × 3 × 1 Monkhorst-Pack k-point mesh were used in this work. The model of large vacancy of NiN2 was set based on the experimental results of pore size (average size, 0.6–0.7 nm). For NiN2 with a large vacancy (NiN2-Nin), we removed two complete benzene rings with four additional carbon atoms and replaced two adjacent carbon atoms with two nitrogen atoms. Meanwhile, for NiN2 with a small vacancy (NiN2-SV), we removed a complete benzene ring and replaced two adjacent carbon atoms with two nitrogen atoms.

The theoretical potentials for NiN2-Nin, NiPc/NC and NiN2-SV were determined, assuming based on a conventional CO2RR mechanism. The elementary steps of the conventional pathway, are believed to involve CO2, COOH, CO adsorbed on the surface (*) according to the following:

$${{{{\rm{CO}}}}}_{2}+* \to {{{{\rm{CO}}}}}_{{2}^{*}}$$
(7)
$${{{{\rm{CO}}}}}_{{2}^{*}}+{{{{\rm{H}}}}}^{+}+{{{{\rm{e}}}}}^{-}\to {{{{\rm{COOH}}}}}^{*}$$
(8)
$${{{{\rm{COOH}}}}}^{*}+{{{{\rm{H}}}}}^{+}+{{{{\rm{e}}}}}^{-}\to {{{{\rm{CO}}}}}^{*}+{{{{\rm{H}}}}}_{2}{{{\rm{O}}}}$$
(9)
$${{{{\rm{CO}}}}}^{*}\to {{{\rm{CO}}}}+\ast$$
(10)

For each step, the Gibbs free energy differences of these intermediates include zero-point energy (ZPE) correction, enthalpic temperature correction, and entropy contribution. Thus, the calculated Gibbs free energy is defined as: ∆G = EDFT + ZPE + ∫CpdT − TdS. The details of ZPE, enthalpic temperature correction, entropy contribution are listed in Supplementary Table S4. The gas values for above three parts correction are sourced from the literature51. ZPE values for the intermediates are calculated from the relevant frequency calculation of VASP, where \({ZPE}={\sum }_{i}\left(\frac{1}{2}*{{hv}}_{i}\right)\). Enthalpic temperature correction and entropy contribution of the intermediates are calculated by vaspkit52. Finite displacement methods are used for computing vibrational modes, with the displacement of each ion set at 0.015 Å. The mean absolute error (MAE) correction of gas-phase molecules for CO2 and CO were +0.41 and −0.18 eV with the RPBE functional51. For the solvation effect, we implemented the implicit solvation model code using VASPsol53,54 after structural optimization, where the corrections of the solvation energy for COOH*, CO*, CO2*, and H* are −0.24, 0.01, −0.12 and 0 eV.

To deal with the hydrogen electrode, the computational hydrogen electrode proposed by Nørskov in 200455 was used. Where, the free energy of proton−electron pair G (H+ + e) = ½ G(H2).

For these four elementary steps:

$${\Delta {{{\rm{G}}}}}_{1}={{{\rm{G}}}}({{{{\rm{CO}}}}}_{{2}^{*}})-{{{{\rm{E}}}}}^{*}-{{{\rm{G}}}}({{{{\rm{CO}}}}}_{2})$$
(11)
$${\Delta {{{\rm{G}}}}}_{2}={{{\rm{G}}}}({{{{\rm{COOH}}}}}^{*})-{{{\rm{G}}}}({{{{\rm{CO}}}}}_{{2}^{*}})-1/2{{{\rm{G}}}}({{{{\rm{H}}}}}_{2})$$
(12)
$${\Delta {{{\rm{G}}}}}_{3}={{{\rm{G}}}}({{{{\rm{CO}}}}}^{*}){{{\mbox{-}}}}{{{\rm{G}}}}({{{{\rm{COOH}}}}}^{*})-1/2{{{\rm{G}}}}({{{{\rm{H}}}}}_{2})+{{{\rm{G}}}}({{{{\rm{H}}}}}_{2}{{{\rm{O}}}})$$
(13)
$${\Delta {{{\rm{G}}}}}_{4}={{{{\rm{E}}}}}^{*}+{{{\rm{G}}}}({{{\rm{CO}}}}){{{\rm{\hbox{-}}}}}{{{\rm{G}}}}({{{{\rm{CO}}}}}^{*})$$
(14)