Auto-tandem CO2 reduction by reconstructed Cu imidazole framework isomers: Unveiling pristine MOF-mediated CO2 activation

Xiang-Da Zhang Jian-Mei Huang Xiaorong Zhu Chang Liu Yue Yin Jia-Yi Huang Yafei Li Zhi-Yuan Gu

Citation:  Xiang-Da Zhang, Jian-Mei Huang, Xiaorong Zhu, Chang Liu, Yue Yin, Jia-Yi Huang, Yafei Li, Zhi-Yuan Gu. Auto-tandem CO2 reduction by reconstructed Cu imidazole framework isomers: Unveiling pristine MOF-mediated CO2 activation[J]. Chinese Chemical Letters, 2025, 36(5): 109937. doi: 10.1016/j.cclet.2024.109937 shu

Auto-tandem CO2 reduction by reconstructed Cu imidazole framework isomers: Unveiling pristine MOF-mediated CO2 activation

English

  • Cu-based metal-organic frameworks (MOFs) have been recognized as promising catalysts in the CO2 reduction reaction (CO2RR) catalysts for generating of C2+ products [17]. It has been widely accepted that Cu-based MOFs usually undergo in-situ reconstruction, yielding Cu-base derivatives, such as Cu, Cu2O, CuO species and their heterostructures [811]. These generated derivatives are identified as paramount active sites for facilitating C—C coupling in CO2RR [11,12]. However, in many cases, distinct MOF precursors can reconstruct the same Cu-based derivatives, but leading to significantly different catalytic performances [1315]. For example, CuII/adeninato/carboxylato MOF (CuII/ade-MOF) nanosheets were converted to Cu nanoparticles, which obtained C2H4 with a maximum faradaic efficiency (FE) of 45% [16]. In another case, similar Cu nanoparticles were derived from semi-conductive MOF-Cu3(HITP)2, where a notable production C2H4 was observed attaining its maximum FE of 63% [17]. The phenomenon indicates that, before the C—C coupling process, there are other activation sites which critically influence catalytic efficiency.

    Acutually, the structures of pristine MOF precursors still remain throughout the catalytic process, particularly in prolonged catalytic cases. In essence, pristine MOFs also harbor the capability to impact the CO2RR process, even though they are commonly overlooked in discussions of the catalytic mechanism. We posit that the varied catalytic outcomes stemming from identical Cu-based derivatives are intricately linked to the presence of pristine MOFs. Unveiling the role of pristine MOFs can not only provide a new direction to design high-proformance MOF-based CO2RR catalysts with minimize trial and error.

    Constructing a rational catalytic system to explore the role of prinstine MOF precursors is still a significant challenge. On the one hand, to avoid variations in the C—C coupling efficiency in the second step, the derivatives produced by different Cu-based MOFs during the catalytic reaction should be completely identical [18,19]. On the other hand, to mitigate the influence of MOF's residual ligands, the prinstine MOFs chosen for comparsion should prossess the same chemical component but differet structures [20]. Last but not lesat, the MOFs catalysts should maintain the srtucture during the catalytic process, forming MOF-derivatives complexes rather than being entirely destroyed [21]. Therefore, to invesitigate a relationship between precursors and catalytic performance, it was essential to ensure structural similarity and stability as well as ligand consistency.

    Here, two Cu(I) imidazole framework isomers with the identical formula [Cu(imidazole)]n were employed as the precursors, namely CuN2 and Cu2N4 based on the coordination status. The only difference between the two precursors was the Cu coordination environments (Scheme 1). In the CuN2 precursor, one Cu atom was linearly coordinated to N atoms of two imidazoles to form zig-zag one-dimensional chain. In the Cu2N4 precursor, two one-dimensional chains was perpendicularly stacked with a Cu-Cu bond formed bewteen the atoms of adjacent layers. The two Cu-based MOFs were both in-situ reconstructed to nanosized Cu(111) with a large number of MOFs retained under cathodic potential, named CuN2/Cu(111) and Cu2N4/Cu(111), respectively. The CO2RR performance of CuN2, Cu2N4, CuN2/Cu(111) and Cu2N4/Cu(111) was evaluated, respectively. The results showed that the both CuN2 and Cu2N4 precursors produced CO while both CuN2/Cu(111) and Cu2N4/Cu(111) produced C2+ products. The CuN2 and Cu2N4 precursors showed a maximum FECO of 43.1% and FECO of 26.2% for the production of CO, respectively. Then, the CuN2/Cu(111) and Cu2N4/Cu(111) exhibited a final FEC of 64.8% and 43.9%, coorespondingly. The FEC was closely related to the FECO of the pristine MOF precursors. We demonstrated a auto-tandem catalysis mechanism to reveal the performance of the Cu(imidazole) system. The pristine Cu(imidazole) played an activiation function to convert CO2 into CO, and Cu(111) catalyzed CO to C2+ products. This auto-tandem catalysis mechanism was confirmed by TPD-CO, in-situ ATR-SEIRAS, and DFT computation. These results conclusively illustrated that pristine CuN2 with more exposed Cu-N sites showed lower CO adsorption energy and generated higher CO substrate concentration, accelerating the formation of C2+ products. This rational mechanism provides a promising way and solution for the rational design of Cu-MOF catalysts.

    Scheme 1

    Scheme 1.  The design of auto-tandem catalysts CuN2/Cu(111) and Cu2N4/Cu(111) and the proposed CO2RR catalytic mechanism.

    The CuN2 and Cu2N4 were prepared with the solvothermal method under different synthetic conditions according to the reported literature method [22,23]. The obtained yellowish and brownish powder was characterized with powder X-ray diffraction (PXRD), X-ray photoelectron spectroscopy (XPS), scanning electron microscopy (SEM), transmission electron microscopy (TEM), and high-resolution transmission electron microscopy (HRTEM). PXRD patterns showed that the synthetic materials were consistent with the simulated CuN2 and Cu2N4, respectively. XPS results revealed that the oxidation state of CuN2 was monovalent, and Cu2N4 exhibited primary monovalent and a little bit of bivalence due to additional Cu-Cu bonding (Fig. S1 in Supporting information). SEM images exhibited that both of them were block material piled up layer by layer, and the size reached the micrometer scale. The morphologies were also confirmed by HRTEM images. Besides, there had no crystal diffraction fringes from HRTEM images. The above characterizations revealed the pristine CuN2 and Cu2N4 were successfully prepared and contained no Cu-based nanoparticles (Fig. S2 in Supporting information).

    According to the previously reported method [24], under an electric field, the Cu-MOF underwent these processes of CO2 adsorption, Cu reduction, Cu-N coordinated bond breaking, Cu atom restructuring, Cu nanoparticle growth, and nanoparticle aggregation. Therefore, we reconstructed CuN2 and Cu2N4 by performing 1-h electrolysis from −0.78 V to −1.28 V vs. RHE. The reconstructed electrocatalysts were named as CuN2/Cu(111) and Cu2N4/Cu(111), respectively. During the reconstruction process, we collected the electrochemical data and analyzed the products. Therein, gaseous products were quantified by an online gas chromatograph (GC), and liquid products were quantified by 1H nuclear magnetic resonance spectroscopy (1H NMR). Meanwhile, formate was detected with a high-performance liquid chromatograph (HPLC) (see details in Supporting information).

    To evaluate the CO2RR performance of CuN2/Cu(111) and Cu2N4/Cu(111), we tested them in both H-cell and flow-cell. On the one hand, in H-cell, we measured linear sweep voltammetry (LSV) after 1-h electrolysis at −1.18 V vs. RHE in both CO2 and N2 saturated 0.1 mol/L KCl electrolyte (Fig. 1a). The CuN2/Cu(111) exhibited lower onset potential and higher total current density (jtotal) than Cu2N4/Cu(111) from −0.4 V to −1.2 V vs. RHE, illustrating better CO2RR catalytic performance. Then, to choose applicable electrolyte, the two catalysts were tested in 0.1 mol/L KHCO3, 0.5 mol/L KHCO3 and 0.1 mol/L KCl electrolyte, respectively (Figs. S3 and S4 in Supporting information). The FE of various products for CuN2/Cu(111) and Cu2N4/Cu(111) were calculated and compared. Obviously, the C2+ product selectivity in 0.1 mol/L KCl was far higher than 0.1 mol/L KHCO3 and 0.5 mol/L KHCO3. Therefore, we chose 0.1 mol/L KCl electrolyte as the test condition in this work. For CuN2/Cu(111), the C2+ products (C2H4, C2H5OH, and n-C3H7OH) were detected from −0.88 V vs. RHE (Figs. S5 and S6 in Supporting information). The maximum FEC reached 64.8%, including FEC of 41.5%, FEC of 18.2%, and FEn-C of 5.1% (Fig. 1b), which was middle level among reported Cu-based materials (Table S1 in Supporting information) [2531]. For Cu2N4/Cu(111), the C2+ products were detected from −0.98 V vs. RHE. The maximum FEC just reached 43.9%, including FEC of 27.1%, FEC of 12.6%, and FEn-C of 3.4% (Fig. 1c). During the operated potential window from −0.98 V to −1.28 V vs. RHE, the FEC2+ of CuN2/Cu(111) was higher than Cu2N4/Cu(111) (Fig. 1d). Besides, to evaluate the number of active sites of CuN2/Cu(111) and Cu2N4/Cu(111), the electrical double-layer capacitor (Cdl) was measured in non-Faraday potential intervals with a different scan rate of cyclic voltammetry (CV) (Fig. S7 in Supporting information) [3235]. The Cdl of CuN2/Cu(111) was calculated as 1.02 mF/cm2, higher than Cu2N4/Cu(111) of 0.32 mF/cm2, indicating more active sites for CO2RR (Fig. 1e). In addition, their stability was tested with online quantification of C2H4 (Fig. 1f). The FEC and jtotal were recorded during 6-h electrolysis. Both FEC and jtotal sharply increased from 5-min to 30-min electrolysis, illustrating the reconstruction of pristine CuN2 and Cu2N4. Then, the FEC2H4 almost maintained, about 40% of CuN2/Cu(111) and 30% of Cu2N4/Cu(111), showing good stability. Cu(imidazole) catalysts exhibit better stability than Cu-O based MOFs during CO2RR process due to the more stable C—N coordination bond based on the hard and soft acid and base (HSAB) theory. According to the above results, it was obvious to conclude that the CO2RR performance of CuN2/Cu(111) was better than Cu2N4/Cu(111).

    Figure 1

    Figure 1.  (a) LSV curves of CuN2/Cu(111) and Cu2N4/Cu(111) in both CO2 and N2 saturated 0.1 mol/L KCl electrolyte. FEs of various products at different operated potentialsl for (b) Cu2N4/Cu(111) and (c) CuN2/Cu(111), respectively. (d) The comparison of FEC2+ for CuN2/Cu(111) and Cu2N4/Cu(111) in H-cell. (e) Cdl measurements with CV method in non-Faraday potential interval under different scan rates. (f) Stability test at −1.18 V vs. RHE for 6-h electrolysis in H-cell. FEs of various products at different operated currents density in flow-cell for (g) Cu2N4/Cu(111), (h) CuN2/Cu(111), respectively. (i) The comparison of FEC2+ for CuN2/Cu(111) and Cu2N4/Cu(111) in flow-cell.

    On the other hand, the CO2RR performance of CuN2/Cu(111) and Cu2N4/Cu(111) was also tested in flow-cell (Figs. 1g-i). The FEC of CuN2/Cu(111) and Cu2N4/Cu(111) were better than H-cell, where maximum was 72.1% and 71.0% at −700 mA/cm2, respectively. Within the operated current range, the performance of CuN2/Cu(111) was better than Cu2N4/Cu(111). The observed phenomenon from flow-cell was similar as in H-cell. Nevertheless, the difference in C2+ performance between the two catalysts less obvious than in H-cell, especially at high current density, this was owing to the destruction of pristine materials from the PXRD patterns (Fig. S12 in Supporting information). Therefore, we subsequently conducted material characterization and mechanism investigation under H-cell conditions with more stable three-phase reaction interface.

    To reveal the actual catalytic sites during the above processes, the structure of CuN2/Cu(111) and Cu2N4/Cu(111) was characterized by PXRD, XPS, Cu LMM Auger spectra, in-situ X-ray absorption near edge structure (XANES), in-situ X-ray emission spectroscopy (XES), HRTEM, and selected area electron diffraction (SAED), respectively (Fig. 2, Figs. S7, S8, S19 and S20 in Supporting information). PXRD patterns showed main diffraction peaks of pristine CuN2 and Cu2N4 were retained (Fig. 2a), which was also observed from SEM images (Fig. S8). Meanwhile, the diffraction peaks of Cu(111) were detected at 43.2°. These results were further confirmed by HRTEM and SAED images (Figs. 2d and e). For CuN2/Cu(111), HRTEM images showed the block of pristine CuN2 piled up layer by layer, and some nanoparticles dispersed in it. The crystal spacing of these nanoparticles was measured as 0.209 nm of Cu(111). In addition, the bright diffraction spots of SAED images also confirmed that Cu(111) was a predominant crystal plane. The selective crystal plane derivatization may be attributed to the uniform Cu(I) distribution in the precursors. Furthermore, the Cu(111) plane was beneficial to promote the formation of C2+ products [3638]. Similarly, the characterization and analysis for Cu2N4/Cu(111) were conducted (Fig. 2e). It showed pristine Cu2N4 was maintained with the formation of Cu(111) nanoparticles. It was worth noting that other very minor Cu species also existed in the two catalysts, such as 0.243 nm of Cu2O(111). Therefore, the derivatives of CuN2 and Cu2N4 were proved to be consistent, at the same time, the pristine materials were retained. Besides, the oxidation states of CuN2/Cu(111) and Cu2N4/Cu(111) were investigated with Cu LMM, in-situ XANES and XES spectra. As shown in Fig. S9 (Supporting information), the main Cu LMM peak of both CuN2/Cu(111) and Cu2N4/Cu(111) was recorded at the kinetic energy of 916.4 eV belonging to Cu(I), demonstrating the reservation of pristine CuN2 and Cu2N4. Meanwhile, the Cu(0) peak with a kinetic energy of 918.8 eV was observed, indicating the reconstruction of Cu(imidazole). More convincingly, compared to pristine CuN2 and Cu2N4, the in-situ XANES spectra of CuN2/Cu(111) and Cu2N4/Cu(111) shifted to a low-energy region during the electrolysis, showing that the oxidation state of Cu(I) was reduced to Cu(0~I) (Fig. 2b). Besides, in-situ XES also exhibited the Cu oxidation state of pristine CuN2 and Cu2N4 were reduced from OCP to −1.3 V vs. RHE (Fig. 2c). In especial, the reconstruction of CuN2 was faster than Cu2N4, and the mixed Cu(0~I) was more than Cu2N4, these were what made the better catalytic performance of derived CuN2/Cu(111). Besides, both in-situ XANES and XES results demonstrated that pristine CuN2 and Cu2N4 did not fully converted to metallic Cu. Therefore, the pristine materials played an important role in these auto-tandem catalysis. Furthermore, to demonstrate the stability of the catalysts, the structure of CuN2/Cu(111) and Cu2N4/Cu(111) after the 6-h stability test was analyzed. PXRD patterns showed the main diffraction peaks of pristine CuN2 and Cu2N4 were still retained (Fig. S10 in Supporting information). The diffraction peak of the Cu(111) derivative was observed, and the intensity was stronger than the materials after 1-h electrolysis, indicating a continuous reconstruction process during the stability test. The Cu(111) crystal plane was also identified from HRTEM and SAED images of CuN2/Cu(111) and Cu2N4/Cu(111) (Fig. S11 in Supporting information).

    Figure 2

    Figure 2.  The characterization of CuN2/Cu(111) and Cu2N4/Cu(111) compared with pristine CuN2 and Cu2N4. (a) PXRD patterns. (b) In-situ XANES spectra. The inset is enlarged part from 8975 eV to 8980 eV. (c) In-situ Cu Kβ1, 3 XES spectra from OCP to −1.3 V vs. RHE. OCP was circuit potential. The microscopic analysis of (d) CuN2/Cu(111) and (e) Cu2N4/Cu(111). The inset is HRTEM and SAED images.

    According to the above structural characterizations, both the pristine Cu MOFs and the same derived Cu(111) existed during the catalytic process. Based on their different catalytic results, it was reasonable to speculate that both pristine Cu MOFs and derived Cu(111) acted as catalysts during the CO2RR process. As Cu(111) usually reported for the C—C coupling to forming C2+ products, we supposed that the pristine Cu MOFs played an activation function to active CO2, forming active substance for the next step of C—C coupling. This catalysis system with two steps was named as the auto-tandem catalysis system. The different catalytic performance of CuN2/Cu(111) and Cu2N4/Cu(111) was possible came from the variation of coordination environment. Particularly, the Cu atom of pristine CuN2 was just coordinated with two N atoms, while for Cu2N4, Cu was not only coordinated with two N atoms but also bonded with the other Cu atom of the adjacent layer. Therefore, we legitimately attributed the different catalytic performances to the different coordination environments of the pristine materials.

    To reveal the role of coordination environment in MOF precursor, we first investigated the performance of sole CuN2 and Cu2N4 by LSV and FE measurements, respectively. The LSV curve of pristine CuN2 also showed lower onset potential and higher jtotal than Cu2N4, indicating better catalytic performance (Fig. 3a). Then, to avoid catalyst reconstruction, short CO2RR experiments with 50-s electrolysis were performed from −0.88 V to −1.18 V vs. RHE for CuN2 and Cu2N4, and respective CO2RR products were quantified. There were no C2+ products, and just CO and H2 was detected for both CuN2 and Cu2N4 (Fig. S13 in Supporting information). For pristine CuN2, CO product was detected at −0.88 V vs. RHE, the FECO increased with potentials, and the maximum FECO reached up to 43.1% (Fig. 3b), which lies a medium level in the Cu-based materials with single metal site (Table S2 in Supporting information). Meanwhile, for pristine Cu2N4, the maximum FECO was only 24.4% (Fig. 3b). Obviously, the CO selectivity of pristine CuN2 was much better than Cu2N4 during the whole operated potential window. These results illustrated the CO2RR performance of pristine CuN2, including onset potential and FECO, was better than Cu2N4.

    Figure 3

    Figure 3.  (a) LSV curves of pristine CuN2 and Cu2N4 in both CO2- and N2-saturated 0.1 mol/L KCl electrolyte. (b) FECO at different potentials with 50-s electrolysis for pristine CuN2 and Cu2N4, respectively. (c) FEco for the pristine materials and FEC2+ for the reconstructed materials at various potentials. (d) After 8-h electrolysis tests for CuN2 and Cu2N4 at −1.18 V vs. RHE, the pristine MOF was reduced to Cu(111), which was immediately tested to obtain FECO. (e) Relative rates of CO2 reduction to CO vs. CO reduction to C2+ products for CuN2 and Cu2N4 after 8-h electrolysis at −1.18 V vs. RHE. (f) CORR of CuN2/Cu(111) and Cu2N4/Cu(111) at various CO volume concentrations with 1-h electrolysis under fixed −1.18 V vs. RHE.

    To confirm the catalyst structure after 50-s electrolysis, the HRTEM, SEM, and PXRD characterizations were executed (Figs. S14-S16 in Supporting information). Both CuN2 and Cu2N4 still exhibited pristine structures, and no Cu-based nanoparticles were observed. Therefore, the above catalytic performance factually reflected the ability of the pristine CuN2 and Cu2N4 to convert CO2. To establish a connection between MOF and Cu(111) in the auto-tandem catalysis system, we made comparison plots of FECO for the pristine materials and FEC for the reconstructed materials at various potentials. As shown in Fig. 3c, both FEC of CuN2/Cu(111) and FEco of CuN2 were higher than Cu2N4/Cu(111) and Cu2N4, respectively. Therefore, the FEC for reconstructed materials exhibited a positive correlation with FECO for the pristine materials.

    According to some literatures, derived Cu metal was also known to convert CO2 into C2+ products through CO as an intermediate. Therefore, in the auto-tandem catalysis system, it is very important to determine which of pristine MOF and the derived Cu(111) is the main active site for converting CO2 to CO. On one hand, pristine CuN2 and Cu2N4 were destructed under 8-h electrolysis tests at −1.18 V vs. RHE, which was fully reduced to Cu(111) from PXRD and SEM analysis (Fig. S17 in Supporting information). Subsequently, the derived Cu(111) was immediately tested to obtain FECO (Fig. 3d). After destruction, the derived Cu(111) exhibited lower FECO than pristine MOF from −0.98 V to −1.18 V vs. RHE. Especially at the potential of C2+ products, pristine CuN2 and Cu2N4 played dominant active sites of converting CO2 to CO. On the other hand, to explore the relationship between the two steps in the auto-tandem catalysis system, we compared the relative rates of CO2 reduction to CO vs. CO reduction to C2+ products for the derived Cu(111) at −1.18 V vs. RHE. (Fig. 3e). All of the rates were around 2 × 10−7 mol min−1 cm−2. The former is not much faster than the latter. Therefore, improving the rate of CO2 reduction to CO may promote C2+ products selectivity due to the rate-determining step of C—C coupling.

    To further investigate the effect of the substrate CO on the C2+ product, control experiments of CO reduction reaction (CORR) with various CO volume concentrations were performed on CuN2/Cu(111) and Cu2N4/Cu(111), respectively. The CO/N2 mixture with CO volume concentrations of 0.2, 0.4, and 1 was employed instead of pure CO2 for electrolysis. At these conditions, the CORR performance of CuN2/Cu(111) and Cu2N4/Cu(111) was evaluated at a potential of −1.18 V vs. RHE. As shown in Fig. 3f, for the CO volume concentration of 0.2, 0.4, and 1, FEC of CuN2/Cu(111) was obtained as 27.3%, 42.9%, and 77.3%, respectively, while FEC of Cu2N4/Cu(111) was 24.3%, 45.6%, and 75.9%, respectively. Therefore, the FEC of CuN2/Cu(111) and Cu2N4/Cu(111) significantly increased with CO volume concentration, illustrating that the CO concentration directly determines the C2+ product formation. Meanwhile, the obtained FEC of CuN2/Cu(111) and Cu2N4/Cu(111) was almost identical under the same CO volume concentration, revealing that CORR bypassed the first CO2 activation step and only depended on the same active site, in this case, the Cu(111). Therefore, these results further demonstrated the rationality of the proposed auto-tandem catalysis mechanism.

    To comprehensively reveal the role of pristine CuN2 and Cu2N4 in the CO2RR process, in-situ attenuated total reflection-surface enhanced infrared reflection absorption spectroscopy (in-situ ATR-SEIRAS) was conducted to monitor the intensity difference of *CO intermediate (Figs. 4a and b, Fig. S20) [3944]. The IR spectrum was collected from open circuit potential (OCP) to −1.28 V vs. RHE. Notably, the *COOH intermediate was monitored at 1652 cm−1 and 1657 cm−1, revealing the activation of CO2. According to the literature [40,45], the band located in the region from 2000 cm−1 to 2100 cm−1 was attributed to the CO linear adsorption on the top sites of Cu surface (*CO), which was the key intermediate for subsequent C—C coupling step of C2+ products formation. On the one hand, for pristine CuN2, the *CO intermediate with a wavenumber of 2057 cm−1 was observed at −0.68 V vs. RHE. Meanwhile, the intensity of *CO intermediate increased with the potentials, indicating faster conversion of CO2 to *CO. The *CO stretching frequency shifted to a lower wavenumber as the potentials decreased from −0.68 V to −1.28 V vs. RHE owing to the Stark effect [4547]. On the other hand, for pristine Cu2N4, the *CO intermediate was observed at a more negative potential of −0.98 V vs. RHE and a slightly lower wavenumber of 2037 cm−1 than CuN2, indicating a higher energy barrier to obtain *CO intermediate. Meanwhile, to quantitatively compare the abilities of Cu sites in CuN2 and Cu2N4 for the conversion of CO2 into *CO, the *CO intermediate band intensities were plotted from −0.08 V to −1.28 V vs. RHE (Fig. 4c). Obviously, the *CO intermediate band intensity of CuN2 was stronger than that of Cu2N4, and it significantly increased with an increase of potentials, which further supported that CuN2 easily produced more *CO intermediates than Cu2N4.

    Figure 4

    Figure 4.  In-situ ATR-SEIRAS results of (a) CuN2 and (b) Cu2N4. On the right were the enlarged spectra from 2090 cm−1 to 2000 cm−1. OCP was circuit potential. (c) *CO band intensity of CuN2 and Cu2N4 from (a) and (b) at different potentials. (d) TPD-CO curve for CuN2 and Cu2N4. (e) DOS calculations of CuN2 and Cu2N4. (f) Gibbs free energy profiles for CO production on CuN2 and Cu2N4. The inset images were the DFT calculation structure model of *COOH adsorbed on Cu site for CuN2 and Cu2N4.

    Besides, the temperature-programmed desorption of CO (TPD-CO) technique was employed to investigate the chemisorption intensity of CO for pristine CuN2 and Cu2N4 (Fig. 4d). Before testing TPD-CO, the thermogravimetric analysis (TGA) was conducted to determine the pyrolysis temperature of the catalysts (Fig. S21 in Supporting information). CuN2 and Cu2N4 showed almost identical TGA curves, indicating that they were both thermally stable before about 250 ℃. Therefore, we collected the TPD-CO data from 50 ℃ to 250 ℃. According to the TPD-CO results, pristine CuN2 exhibited a much lower CO desorption temperature of 148 ℃ than 207 ℃ of Cu2N4, indicating a lower desorption activation energy, which facilitated the *CO desorption and subsequent C—C coupling step on Cu(111) site.

    Furthermore, to understand the different adsorption strength for intermediates during the CO2RR process, we analyzed the density of state (DOS) for CuN2 and Cu2N4. Especially, Cu 3d orbital partial density of state (PDOS) was conducted to calculate the d-band center which was regarded as the applicable descriptor to explain catalytic reactivity [48,49]. As shown in Fig. 4e, d-center value of CuN2 and Cu2N4 were −1.4 eV and −2.9 eV, respectively. The position of d-band center for CuN2 was closer to the Femi level (Ef) than Cu2N4, indicating stronger adsorption of intermediates [50]. Besides, the reaction pathway and energy barrier of *CO intermediate formation were explored with the DFT calculation [5153]. The Cu sites of pristine CuN2 and Cu2N4 were modeled as the active sites for CO production. The whole reaction pathway from CO2 to CO was proposed as follows (Fig. 4f) [44]: CO2 → *COOH → *CO → CO, where the first step CO2 → *COOH was the rate-determining step (RDS). Meanwhile, the intermediates such as *COOH and *CO were also detected with in-situ ATR-SEIRAS spectroscopy. The RDS reaction energy of pristine CuN2 was calculated as 0.46 eV, which was lower than pristine Cu2N4 with 0.57 eV. These results further demonstrated that the pristine coordination environment of Cu(imidazole) catalysts played a decisive role in CO generation, thus affecting further C2+ production.

    In summary, we uncovered that the pristine Cu(imidazole) played an critical activation function to convert CO2 into CO, and subsequent the in-situ derived Cu(111) acted as the active sites for C—C coupling. This auto-tandem catalysis activity was closely related to the coordination environment of the pristine MOFs. CuN2 with more exposed Cu-N sites showed higher ability to form CO, thus the CuN2/Cu(111) exhibited higher C2+ selectivity. The catalytic active sites were explored with HRTEM, SAED, XPS, and XANES techniques. This mechanism was demonstrated with CORR control experiment, TPD-CO, in-situ ATR-SEIRAS, and DFT calculation. We anticipate that the auto-tandem catalysis mechanism will be utilized in the design and development of Cu-based MOF precursors for promoting the C2+ products selectivity of reconstructed catalysts. The key of applying the auto-tandem catalytic mechanism is to improve the intrinsic catalytic active and stability of pristine MOF.

    Xiang-Da Zhang: Conceptualization, Data curation, Formal analysis, Writing – original draft. Jian-Mei Huang: Data curation. Xiaorong Zhu: Software. Chang Liu: Investigation. Yue Yin: Investigation. Jia-Yi Huang: Investigation. Yafei Li: Supervision, Validation. Zhi-Yuan Gu: Funding acquisition, Supervision, Visualization, Writing – review & editing.

    This work is supported by the National Natural Science Foundation of China (Nos. 22174067 and 22204078), the Natural Science Foundation of Jiangsu Province of China (No. BK20220370), Jiangsu Provincial Department of Education (No. 22KJB150009), State Key Laboratory of Analytical Chemistry for Life Science (No. SKLACLS2218), and the Priority Academic Program Development of Jiangsu Higher Education Institutions. The authors would like to thank BL11B, BL14W1, and BL17B in Shanghai Synchrotron Radiation Facility (SSRF) and BL01B in National Synchrotron Radiation Laboratory (NSRL).

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2024.109937.


    1. [1]

      J. Liu, D. Yang, Y. Zhou, et al., Angew. Chem. Int. Ed. 60 (2021) 14473–14479. doi: 10.1002/anie.202103398

    2. [2]

      D.H. Nam, O. Shekhah, G. Lee, et al., J. Am. Chem. Soc. 142 (2020) 21513–21521. doi: 10.1021/jacs.0c10774

    3. [3]

      X.F. Qiu, H.L. Zhu, J.R. Huang, et al., J. Am. Chem. Soc. 143 (2021) 7242–7246. doi: 10.1021/jacs.1c01466

    4. [4]

      J.X. Wu, S.Z. Hou, X.D. Zhang, et al., Chem. Sci. 10 (2019) 2199–2205. doi: 10.1039/c8sc04344b

    5. [5]

      A.M. Appel, J.E. Bercaw, A.B. Bocarsly, et al., Chem. Rev. 113 (2013) 6621–6658. doi: 10.1021/cr300463y

    6. [6]

      S. Nitopi, E. Bertheussen, S.B. Scott, et al., Chem. Rev. 119 (2019) 7610–7672. doi: 10.1021/acs.chemrev.8b00705

    7. [7]

      L. Majidi, A. Ahmadiparidari, N. Shan, et al., Adv. Mater. 33 (2021) 2004393.

    8. [8]

      T. Ahmad, S. Liu, M. Sajid, et al., Nano Res. Energy 1 (2022) e9120021. doi: 10.26599/nre.2022.9120021

    9. [9]

      X. Han, Z. Liu, M. Cao, et al., Chem. Mater. 34 (2022) 6713–6722. doi: 10.1021/acs.chemmater.2c00444

    10. [10]

      C. Kong, G. Jiang, Y. Sheng, et al., Chem. Eng. J. 460 (2023) 141803.

    11. [11]

      J. Wang, Y. Zhang, Y. Ma, et al., ACS Mater. Lett. 4 (2022) 2058–2079. doi: 10.1021/acsmaterialslett.2c00751

    12. [12]

      X. Tan, C. Yu, C. Zhao, et al., ACS Appl. Mater. Interfaces 11 (2019) 9904–9910. doi: 10.1021/acsami.8b19111

    13. [13]

      M.K. Kim, H.J. Kim, H. Lim, et al., Electrochim. Acta 306 (2019) 28–34. doi: 10.13000/jfmse.2019.2.31.1.28

    14. [14]

      M.R. Smith, A. Gilman, C.W. Hullfish, et al., J. Phys. Chem. C 126 (2022) 13649–13659. doi: 10.1021/acs.jpcc.2c01287

    15. [15]

      Z. Weng, Y. Wu, M. Wang, et al., Nat. Commun. 9 (2018) 415.

    16. [16]

      F. Yang, A. Chen, P.L. Deng, et al., Chem. Sci. 10 (2019) 7975–7981. doi: 10.1039/c9sc02605c

    17. [17]

      H. Sun, L. Chen, L. Xiong, et al., Nat. Commun. 12 (2021) 6823.

    18. [18]

      Y. Zhao, L. Zheng, D. Jiang, et al., Small 17 (2021) 2006590.

    19. [19]

      Q. Zhu, D. Yang, H. Liu, et al., Angew. Chem. Int. Ed. 59 (2020) 8896–8901. doi: 10.1002/anie.202001216

    20. [20]

      N. Li, P. Yan, Y. Tang, et al., Appl. Catal. B 297 (2021) 120481.

    21. [21]

      D. Yao, T. Cheng, A. Vasileff, et al., Angew. Chem. Int. Ed. 60 (2021) 18178–18184. doi: 10.1002/anie.202104747

    22. [22]

      Y.Q. Tian, H.J. Xu, L.H. Weng, et al., Eur. J. Inorg. Chem. 2004 (2004) 1813–1816.

    23. [23]

      X.C. Huang, J.P. Zhang, Y.Y. Lin, et al., Chem. Commun. (2004) 1100–1101.

    24. [24]

      T. Bao, W. Zhai, C. Yu, et al., Small Struct. 5 (2024) 2300293.

    25. [25]

      Y. Li, Z. Pei, D. Luan, et al., Angew. Chem. Int. Ed. 62 (2023) e202302128.

    26. [26]

      Y. Baek, H. Song, D. Hong, et al., J. Mater. Chem. A 10 (2022) 9393–9401. doi: 10.1039/d1ta10345h

    27. [27]

      Z. Li, Y. Fang, J. Zhang, et al., J. Mater. Chem. A 9 (2021) 19932–19939. doi: 10.1039/d1ta02565a

    28. [28]

      L. Fan, C. Xia, F. Yang, et al., Sci. Adv. 6 (2020) eaay3111.

    29. [29]

      A. Handoko, C. Ong, Y. Huang, et al., J. Phys. Chem. C 120 (2016) 20058–20067. doi: 10.1021/acs.jpcc.6b07128

    30. [30]

      A. Loiudice, P. Lobaccaro, E.A. Kamali, et al., Angew. Chem. Int. Ed. 55 (2016) 5789–5792. doi: 10.1002/anie.201601582

    31. [31]

      D. Kim, C.S. Kley, Y. Li, et al., Proc. Natl. Acad. Sci. U. S. A. 114 (2017) 10560–10565. doi: 10.1073/pnas.1711493114

    32. [32]

      Y. Sha, J. Zhang, X. Cheng, et al., Angew. Chem. Int. Ed. 61 (2022) e202200039.

    33. [33]

      J. Yuan, S. Chen, Y. Zhang, et al., Adv. Mater. 34 (2022) 2203139.

    34. [34]

      Y. Li, S.L. Zhang, W. Cheng, et al., Adv. Mater. 34 (2022) 2105204.

    35. [35]

      Y. Zou, C. Liu, C. Zhang, et al., Nat. Commun. 14 (2023) 5780.

    36. [36]

      T. Yan, P. Wang, W.Y. Sun, Small 19 (2023) 2206070.

    37. [37]

      H. Xiao, T. Cheng, W.A. Goddard Ⅲ, J. Am. Chem. Soc. 139 (2017) 130–136. doi: 10.1021/jacs.6b06846

    38. [38]

      Z. -H. Zhao, K. Zheng, N. -Y. Huang, et al., Chem. Commun. 57 (2021) 12764–12767. doi: 10.1039/d1cc05376k

    39. [39]

      B. r. Ratschmeier, B. r. Braunschweig, J. Phys. Chem. C 125 (2021) 16498–16507. doi: 10.1021/acs.jpcc.1c02898

    40. [40]

      J. Heyes, M. Dunwell, B. Xu, J. Phys. Chem. C 120 (2016) 17334–17341. doi: 10.1021/acs.jpcc.6b03065

    41. [41]

      H. Wang, Y.W. Zhou, W.B. Cai, Curr. Opin. Electrochem. 1 (2017) 73–79.

    42. [42]

      H. Guo, D.H. Si, H.J. Zhu, et al., Angew. Chem. Int. Ed. 63 (2024) e202319472.

    43. [43]

      Q.J. Wu, D.H. Si, Q. Wu, et al., Angew. Chem. Int. Ed. 62 (2023) e202215687.

    44. [44]

      H.J. Zhu, D.H. Si, H. Guo, et al., Nat. Commun. 15 (2024) 1479.

    45. [45]

      H. Miyake, M. Osawa, Chem. Lett. 33 (2004) 278–279.

    46. [46]

      D.K. Lambert, Electrochim. Acta 41 (1996) 623–630.

    47. [47]

      A. Wuttig, C. Liu, Q. Peng, et al., ACS Central. Sci. 2 (2016) 522–528. doi: 10.1021/acscentsci.6b00155

    48. [48]

      S. Jiao, X. Fu, H. Huang, Adv. Funct. Mater. 32 (2022) 2107651.

    49. [49]

      E. Santos, W. Schmickler, ChemPhysChem 7 (2006) 2282–2285. doi: 10.1002/cphc.200600441

    50. [50]

      Q. Zhang, L. Guo, J. Clust. Sci. 29 (2018) 867–877. doi: 10.1007/s10876-018-1346-x

    51. [51]

      G. Kresse, J. Hafner, Phys. Rev. B 47 (1993) 558–561.

    52. [52]

      K. Kim, P. Wagner, K. Wagner, et al., Energy Fuel 36 (2022) 4653–4676. doi: 10.1021/acs.energyfuels.2c00400

    53. [53]

      J. Wang, L. Gan, Q. Zhang, et al., Adv. Energy Mater. 9 (2019) 1803151.

  • Scheme 1  The design of auto-tandem catalysts CuN2/Cu(111) and Cu2N4/Cu(111) and the proposed CO2RR catalytic mechanism.

    Figure 1  (a) LSV curves of CuN2/Cu(111) and Cu2N4/Cu(111) in both CO2 and N2 saturated 0.1 mol/L KCl electrolyte. FEs of various products at different operated potentialsl for (b) Cu2N4/Cu(111) and (c) CuN2/Cu(111), respectively. (d) The comparison of FEC2+ for CuN2/Cu(111) and Cu2N4/Cu(111) in H-cell. (e) Cdl measurements with CV method in non-Faraday potential interval under different scan rates. (f) Stability test at −1.18 V vs. RHE for 6-h electrolysis in H-cell. FEs of various products at different operated currents density in flow-cell for (g) Cu2N4/Cu(111), (h) CuN2/Cu(111), respectively. (i) The comparison of FEC2+ for CuN2/Cu(111) and Cu2N4/Cu(111) in flow-cell.

    Figure 2  The characterization of CuN2/Cu(111) and Cu2N4/Cu(111) compared with pristine CuN2 and Cu2N4. (a) PXRD patterns. (b) In-situ XANES spectra. The inset is enlarged part from 8975 eV to 8980 eV. (c) In-situ Cu Kβ1, 3 XES spectra from OCP to −1.3 V vs. RHE. OCP was circuit potential. The microscopic analysis of (d) CuN2/Cu(111) and (e) Cu2N4/Cu(111). The inset is HRTEM and SAED images.

    Figure 3  (a) LSV curves of pristine CuN2 and Cu2N4 in both CO2- and N2-saturated 0.1 mol/L KCl electrolyte. (b) FECO at different potentials with 50-s electrolysis for pristine CuN2 and Cu2N4, respectively. (c) FEco for the pristine materials and FEC2+ for the reconstructed materials at various potentials. (d) After 8-h electrolysis tests for CuN2 and Cu2N4 at −1.18 V vs. RHE, the pristine MOF was reduced to Cu(111), which was immediately tested to obtain FECO. (e) Relative rates of CO2 reduction to CO vs. CO reduction to C2+ products for CuN2 and Cu2N4 after 8-h electrolysis at −1.18 V vs. RHE. (f) CORR of CuN2/Cu(111) and Cu2N4/Cu(111) at various CO volume concentrations with 1-h electrolysis under fixed −1.18 V vs. RHE.

    Figure 4  In-situ ATR-SEIRAS results of (a) CuN2 and (b) Cu2N4. On the right were the enlarged spectra from 2090 cm−1 to 2000 cm−1. OCP was circuit potential. (c) *CO band intensity of CuN2 and Cu2N4 from (a) and (b) at different potentials. (d) TPD-CO curve for CuN2 and Cu2N4. (e) DOS calculations of CuN2 and Cu2N4. (f) Gibbs free energy profiles for CO production on CuN2 and Cu2N4. The inset images were the DFT calculation structure model of *COOH adsorbed on Cu site for CuN2 and Cu2N4.

  • 加载中
计量
  • PDF下载量:  0
  • 文章访问数:  165
  • HTML全文浏览量:  9
文章相关
  • 发布日期:  2025-05-15
  • 收稿日期:  2024-03-08
  • 接受日期:  2024-04-28
  • 修回日期:  2024-04-02
  • 网络出版日期:  2024-04-29
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章