Jingui
Zheng
,
Shaohan
Xu
,
Lingzhi
Sun
,
Xun
Pan
,
Qihao
Xie
and
Guohua
Zhao
*
School of Chemical Science and Engineering, Key Laboratory of Spine and Spinal Cord Injury Repair and Regeneration, Ministry of Education, Tongji Hospital, Tongji University, Shanghai 200092, People's Republic of China
First published on 24th February 2025
The electrocatalytic reduction of CO2 and dimethylamine (HN(CH3)2) for C–N coupling is a promising strategy for synthesizing N,N-dimethylformamide (DMF). However, the generation of suitable coupling intermediates via the dehydrogenation of HN(CH3)2 and reduction hydrogenation of CO2 is the key challenge for achieving C–N bonding to synthesize DMF. We optimized the electronic structure of HN(CH3)2 and CO2 for adsorption on indium single atoms by adjusting the coordination structure, thereby promoting the hydrogen transfer from nitrogen of dimethylamine to CO2, which generated the intermediates of *N(CH3)2 and *COOH for C–N bonding to synthesize DMF. The yield of DMF synthesized on InN3 reached 41.3 μmol L−1 h−1, which was about 12 times greater than that of InN4 at −0.8 V. In situ technology and DFT calculations jointly demonstrated that compared with InN4, InN3 optimized the electron distribution of adsorbed CO2 and HN(CH3)2. The electron density of hydrogen on HN(CH3)2 decreased, exhibiting its electrophilic properties. In addition, oxygen of CO2 accumulated electrons near the dimethylamine end and exhibited strong electron-rich properties, which led to hydrogen transfer from dimethylamine to CO2, generating the species *N(CH3)2 and *COOH that are conducive to C–N coupling to synthesize DMF on InN3. This work provides important theoretical guidance for the C–N coupling of CO2 and amines.
Broader contextEnvironmental pollution and energy crisis are considered two major problems that threaten the survival and development of human beings. CO2, which is a typical greenhouse gas, and dimethylamine, which is a water pollutant, need to be removed urgently. The simultaneous conversion of CO2 and dimethylamine via C–N coupling is a promising strategy for synthesizing N,N-dimethylformamide (DMF), which has a wide range of industrial and other applications. However, generation of suitable coupling intermediates via the dehydrogenation of HN(CH3)2 and reduction hydrogenation of CO2 is the key challenge in C–N coupling for synthesizing DMF. In this work, we optimized the electronic structure of HN(CH3)2 and CO2 for adsorption on indium single atoms by adjusting the coordination structure, thereby promoting the hydrogen transfer from dimethylamine to CO2, which generated the intermediates *N(CH3)2 and *COOH for C–N coupling to synthesize DMF. This work provides important theoretical guidance for the C–N coupling of CO2 and secondary amines. |
The electronic structure of the catalyst directly affects the C–N coupling4,5 between ammonia dehydrogenation and CO2 hydrogenation. This is because coordination environment of interface atoms affects the adsorption structure of reactants on the catalyst, thereby altering the electronic structure of adsorbed reactants and improving the adsorption behaviour of the catalyst towards reactants or reaction intermediates,6,7 which optimize the reaction pathway. Research has found that optimizing the electronic structure of catalysts by adjusting the coordination structure can promote the dehydrogenation of ammonia. Zan et al. constructed asymmetric Ru1N2O1 by changing the coordination environment of Ru single-atom catalysts, which not only enhanced its adsorption of ammonia but also optimized the electronic configuration of Ru.8 It led to direct activation of N–H bonds without the involvement of surface oxygen, thereby facilitating dehydrogenation. Moreover, adjusting the electronic structure of the catalyst through the coordination structure can promote the activation of reduction of carbon dioxide. Zhong et al. reported that the electronic structure of nickel was regulated by adjusting the coordination environment around the bimetallic nickel sites. The high electrocatalytic activity of Ni2–N3C4 towards reduction of CO2 could be attributed to the modulation of its electronic structure on the Ni sites and the resulting appropriate binding energy for *COOH and *CO intermediates.9 Besides, owing to the unique electronic properties of indium species, it can be highly selectively reduced to CO2 as HCOOH. It can be seen that formic acid contains HCO groups, and the reduction of CO2 to HCO groups is also an important step in the N-formylation of dimethylamine and CO2 to synthesize DMF.
Therefore, we constructed an indium single-atom catalyst with a low coordination structure, which optimized its electronic structure via adjusting the coordination environment for the reduction of CO2 and dimethylamine to synthesize DMF. InN3 with low N-coordinated indium single-atom catalyst led to easy dehydrogenation of dimethylamine, which was then coupled with the CO2 reduction intermediate for C–N-coupled N-formylation to synthesize DMF. The low coordination structure of indium single atoms significantly affects the electronic properties of the catalyst and reactants. Differential charge analysis revealed that the indium single-atom catalyst of InN3 affects the electron distribution of its adsorbed reactants dimethylamine and CO2. The electron density of hydrogen of nitrogen on dimethylamine decreases, exhibiting electrophilic properties. In addition, the oxygen atom of electrons on CO2 accumulates near the dimethylamine end, exhibiting strong electron-rich properties, which weakens the hydrogen atom binding ability of dimethylamine on its nitrogen, thus facilitating dehydrogenation transfer to the oxygen of CO2, achieving hydrogen transfer from dimethylamine to CO2, generating *N(CH3)2 and *COOH intermediates that are conducive to coupling, thereby efficiently forming C–N bonds, and the free energy of N-formylation on InN3 is lower, making it easier to generate DMF. We designed and prepared a low N-coordinated indium single-atom catalyst of InN3 to promote efficient C–N coupling between CO2 and dimethylamine for the electrochemical synthesis of DMF. The yield of DMF synthesized by InN3, a low N-coordinated indium single-atom catalyst, reached 41.3 μmol L−1 h−1 at −0.8 V, with a corresponding faradaic efficiency of 22.4%. Synchrotron radiation, in situ infrared spectroscopy, in situ Raman spectroscopy, and DFT theoretical calculations jointly demonstrate that by regulating the coordination structure of indium single atoms, changes in electronic structure are induced, which promote hydrogen transfer from dimethylamine to CO2. It is more conducive to CO2 reduction and hydrogenation. This efficiently generates intermediates of *N(CH3)2 and *COOH that facilitate C–N coupling to synthesise DMF. This work provides important theoretical guidance for the C–N coupled N-formylation of CO2 and amines, thereby paving the way for advancing electrochemical CO2 conversion technologies.
We used differential charge density to analyse the reason why InN3 achieves catalytic reduction of CO2 and dimethylamine to synthesize DMF. We compared the differential charge density of indium single-atom catalysts (InN4, InN3) with different nitrogen coordination environments, and explained their catalytic mechanism from the aspects of electron distribution. The most critical step in the catalytic reduction of CO2 and dimethylamine to synthesize DMF is the hydrogen transfer between dimethylamine and CO2, generating reaction intermediates *N(CH3) and *COOH that are favourable for C–N coupling. This step involves the dehydrogenation of dimethylamine, which is very difficult to occur in the reduction reaction. We optimized the adsorption configurations of InN4 and InN3 for both CO2 and dimethylamine via differential charge density. The analysis results showed that InN3 adsorbs dimethylamine, and CO2 molecules are perpendicular to the dimethylamine molecules adsorbed on InN3 (Fig. S1, ESI†). Charge transfer occurs on the CO2 molecules, and the electron density of oxygen on CO2 near dimethylamine increases in InN3, exhibiting electron-rich properties. However, the electron density of the hydrogen atom on nitrogen of dimethylamine is missing, resulting in strong electrophilicity (Fig. 1c), which easily causes hydrogen on dimethylamine to detach from dimethylamine and form a bond with the electron-rich oxygen on CO2, leading to hydrogen transfer and the key step of *(CH3)2NH + CO2 → *N(CH3)2 + *COOH. When the InN4 catalyst simultaneously adsorbs CO2 and dimethylamine, CO2 is parallel to the dimethylamine molecules adsorbed on InN4. From the configuration diagram of differential charge density, it can be seen that the side of CO2 near the dimethylamine is electron deficient, and the position of hydrogen on the primary amine of the dimethylamine molecule is also electron deficient, which causes charge repulsion. Therefore, hydrogen transfer between dimethylamine molecule and CO2 molecule is not easy to occur on InN4, completing the key step of *(CH3)2NH + CO2 → *N(CH3)2 + *COOH. Therefore, the free energy of the key reaction intermediate *N(CH3)2 + *COOH, which is conducive to coupling, generated on InN4 is as high as 0.74 eV, which is twice as high as that on InN3. In summary, the low coordination of InN3 alters the electron arrangement of adsorbed dimethylamine and carbon dioxide, making dimethylamine more prone to dehydrogenation and CO2 reduction hydrogenation, thereby causing C–N coupling between *N(CH3)2 and *COOH intermediates to form DMF. In order to deeply understand the interaction between the hydrogen atom of the amino group in HN(CH3)2 and the oxygen atom of CO2 on InN3, the projected density of states (PDOS) was calculated (Fig. S2, ESI†). The results indicated that the overlap between the s orbital of the hydrogen atom on amino group in HN(CH3)2 and the p orbital of the oxygen atom in CO2 was high, indicating that there was electronic interaction and strong binding force between the hydrogen atom of amino group in HN(CH3)2 and the oxygen atom of CO2, facilitating hydrogen transfer of HN(CH3)2 and CO2 to synthesize DMF.
Synthesis and structural characterization of indium single-atom catalysts. The indium single-atom catalysts with different nitrogen coordinations (InN4 and InN3) were prepared by controlling different calcination temperatures (900 °C, 1000 °C) (Fig. S3, ESI†). From its XRD pattern (Fig. 2a), it can be seen that the material after high-temperature calcination only has a peak shape of carbon and no crystal shape of metal indium nanoparticles. It is preliminarily determined that the catalyst we prepared is not a nanoparticle, possibly because the prepared catalyst is atomically dispersed and does not have its peak structure.10 The N vacancies were confirmed through electron paramagnetic resonance (EPR) testing. All samples exhibit a Lorentz line with a corresponding g-value of 2.003,11 which can be attributed to unpaired electrons in the aromatic ring of localized π-state carbon atoms in the calcined MOF material.12,13 As the calcination temperature increases, the EPR signal of the sample gradually increases, indicating that nitrogen atoms were lost, leaving behind excess electrons. The material of indium-doped ZIF-8 calcined at 1000 °C has the strongest signal, indicating the highest content of N vacancies, possibly InN3. The weakest and almost non-existent signal was obtained by calcination at 900 °C, which may be InN4 (Fig. 2c). Then we demonstrate the specific coordination configuration through synchrotron radiation. By scanning electron microscopy, it can be observed that the prepared indium single-atom InN3 exhibited a regular dodecahedral structure (Fig. S4a and b, ESI†), which is similar in morphology to similar materials prepared in the literature.14–16 Through EDS mapping, it can be observed that In and N atoms are uniformly distributed (Fig. S4c and d, ESI†). The surface chemical properties of the prepared InN3 catalyst was analysed by XPS. We conducted a detailed analysis of the valence states of N and In elements in the catalyst, and N 1s can be fitted into four main components, corresponding to pyridine N (398.4 eV), pyrrole N (399.0 eV), graphite N (401.2 eV), and In–N (400.2 eV)17 (Fig. 2b). We continued to analyze the valence state of indium element in catalysts InN4 and InN3. In 3d5/2 and In 3d3/2 are located at 445.0 eV and 452.5 eV,10,18–20 respectively, indicating that the valence state of In in the prepared catalyst InN3 is +3 (Fig. S5, ESI†). Its valence state is the same as that of the indium species in InN4. We used synchrotron radiation to confirm that the catalyst is a single atom and determine the coordination environment of catalyst. By processing the R space of InN3 and InN4 catalysts prepared by synchrotron radiation detection, the different coordination environments of indium single atoms were confirmed. We used synchrotron X-ray absorption spectroscopy (XAS) to conduct in-depth research on the electronic and coordination structures of the catalyst. By comparing the white line peaks of the k-edge X-ray absorption energy near-edge structure spectra of different indium species (In foil, In2O3, InN3, and InN4), it can be seen that the white line peak positions between InN3 and In2O3 are almost the same, and the valence state of In on InN3 is the same as the oxdation state of In2O3, which is +3 (Fig. 2d). Single-atom catalysts of InN3 and InN4 were confirmed via synchrotron radiation. In Fig. 2e, the k-edge Fourier transform k3-weighted EXAFS spectra (R space) of indium in InN3 and InN4 show only one distinct peak, which is an important indicator of single-atom catalysts.21–23 Compared to In foil, the In2O3 material has In–In bonds, while InN3 and InN4 only have In–N bonds at 1.53 Å, indicating that the material exists in the form of In with N coordination24 (Fig. 2e). We further demonstrated through wavelet transform of synchrotron radiation that the catalyst we prepared is a single-atom catalyst of indium. By comparing the wavelet transform plots of k-edge EXAFS signals of In foil, In2O3, InN3, and InN4, it can be found that there are only In–In bonds in the In foil (Fig. 2j), while In2O3 has both In–In and In–O bonds (Fig. 2k), but there are no In–In bonds in InN3 (Fig. 2m) and InN4 (Fig. 2l), there is only In–N bonds. In addition, the maximum value of k in InN3 is slightly smaller than that in InN4, which may be caused by the low coordinated structure of nitrogen. We further analyzed the coordination configuration of indium single atoms using synchrotron radiation and elaborated on the coordination environment of single-atom indium. By analyzing the R space of InN3 and InN4 (Fig. 2g), it was found that the decrease in peak intensity of In–N bonds indicates a decrease in their coordination number, which roughly proves the successful preparation of single-atom catalysts for InN3 and InN4. In addition, it can be seen from the changes in q-space (Fig. 2f) that our data quality is very good. The coordination number of indium single atoms was determined by fitting indium single-atom catalysts calcined at different temperatures (900 °C, 1000 °C). Through EXAFS fitting of the K-edge of indium, analysis shows that the indium single atom obtained by calcination at 900 °C is connected to four N atoms in the first coordination shell of indium, resulting in indium nitrogen-tetracoordinated indium single-atom InN4. The configuration diagram is shown in Fig. 2i. The indium single atom obtained by calcination at 1000 °C is connected to three N atoms in the first coordination shell of indium, resulting in indium-nitrogen triple-coordinated indium single-atom InN3. The configuration diagram is shown in Fig. 2h. Through synchrotron radiation analysis of the catalysts, it was demonstrated that the prepared catalysts InN3 and InN4 are single-atom catalysts. In addition, it is consistent with the target catalyst model we designed through DFT calculations.
Performance evaluation of InN3 and InN4 for electrocatalytic synthesis of DMF from CO2 and dimethylamine. We conducted performance evaluation on the electrocatalytic reduction of CO2 and dimethylamine to synthesise DMF on InN4 and InN3 single-atom catalysts. Through the linear scan curve test (Fig. 3a), it can be seen that the current of the InN3 single-atom catalyst is higher than that of the InN4 single-atom catalyst, indicating that the InN3 single-atom catalyst is beneficial for catalyzing the reduction of CO2 and dimethylamine. We found that the InN3 catalyst prepared at a calcination temperature of up to 1000 °C exhibited smaller electrochemical impedance than that of InN4 in the electrochemical impedance spectra, making it easier to conduct electrons for electrochemical reduction reactions (Fig. 3b). We determined the double-layer capacitance of InN4 and InN3 single-atom catalysts by testing their cyclic voltammetry spectra at different scan rates in the Faraday interval (Fig. S6a and b, ESI†), and their electrochemical active area was further determined. The slope analysis indicated that the InN3 single atom has a higher electrochemical active area (Fig. 3c). The performance of electrochemical synthesis of DMF from CO2 and dimethylamine was evaluated at different potentials. The yields of N,N-dimethylformamide (DMF) were detected by HPLC (Fig. S7, ESI†). The prepared InN4 and InN3 catalysts were used as working electrodes. A platinum plate was used as the counter electrode, and a saturated calomel electrode was used as the reference electrode to test the performance of DMF synthesis at different voltages. It was found that the InN3 catalyst exhibited the highest electrocatalytic reduction of CO2 and dimethylamine to DMF at −0.8 V, reaching 41.3 μmol L−1 h−1 (Fig. 3g), with a corresponding faradaic efficiency of up to 22.4% (Fig. 3h). However, it can be found that different nitrogen-coordinated indium single-atom catalysts can electrochemically synthesize DMF, but their performance varies greatly. At −0.8 V, the catalyst of InN4 only achieved a yield of 3.4 μmol L−1 h−1 for electrocatalytic reduction of CO2 and dimethylamine to synthesize DMF (Fig. 3d), and the corresponding faradaic efficiency was only 2.4% (Fig. 3e). By comparing the yield of different nitrogen-coordinated indium single-atom catalysts for DMF synthesis at the same optimal potential, it was found that the yield of InN3 electrocatalytic reduction of CO2 and dimethylamine to synthesize DMF at the optimal potential of −0.8 V was about 12 times that of InN4 (Fig. 3f). Comparing the efficiency of reduction of CO2 and dimethylamine to synthesize DMF on InN4, the faradaic efficiency of InN3 in catalyzing DMF synthesis at the optimal potential of −0.8 V is more than 9 times that of InN4 (Fig. 3i). In summary, the unique low coordination structure of InN3 single-atom catalysts exhibits superior catalytic activity for the reduction of carbon dioxide and dimethylamine to synthesize DMF. We also compared the performance of different electrolyte solutions in the synthesis of DMF from CO2 and dimethylamine. The FE for synthesizing DMF in an electrolyte solution of 0.1 M KHCO3 is 22.4%, while the FE in KOH and Na2SO4 is 1.2% and 5.8%, respectively (Fig. S8, ESI†), which are much lower than the efficiency of synthesizing DMF in an electrolyte solution of KHCO3. For the long-term stability test, electrochemical synthesis of DMF was carried out for 12 hours. There is no significant decrease in current density, which indicated that InN3 is relatively stable (Fig. S9, ESI†).
In situ characterization of the mechanism for synthesizing DMF. We investigated the reaction mechanism of different nitrogen-coordinated indium single-atom catalysts (InN4 and InN3) for the catalytic reduction of CO2 and dimethylamine C–N bonding to synthesize DMF via in situ infrared spectroscopy and in situ mass spectrometry. The intermediates involved in the catalytic reduction of CO2 and dimethylamine using indium single-atom catalysts InN3 and InN4 with different nitrogen coordinations were detected by in situ infrared spectroscopy. It combines with intermediate species detected by in situ mass spectrometry to elucidate the reaction mechanism for synthesizing DMF. We characterized the adsorption of CO2 and dimethylamine by InN3 single-atom catalysts via in situ infrared spectroscopy. We found that a large absorption peak at a wavelength of 1241 cm−1 was detected on InN3, and the intensity of the peak gradually increases over time until it no longer changes at 30 minutes, indicating that the catalyst adsorption has reached saturation. The absorption peak at 1241 cm−1 corresponds to the adsorption of dimethylamine25 (Fig. 4a), while InN4 did not detect any adsorption peak for dimethylamine. The adsorption energies of InN3 and InN4 single-atom catalysts for dimethylamine were compared through DFT calculations. Through DFT theoretical calculations, it was found that the adsorption energy of InN3 for dimethylamine is −0.09 eV, while the adsorption energy of InN4 for dimethylamine is 0.64 eV (Fig. 4d). The more negative the free energy, the more likely it is to occur. Therefore, InN3 is more conducive to adsorbing dimethylamine. It has been demonstrated through infrared spectroscopy and DFT calculations that InN3 is more favourable for adsorbing dimethylamine.
The reaction intermediates in the synthesis of DMF from CO2 and dimethylamine in InN3 and InN4 catalysts through in situ infrared spectroscopy were detected to explore the mechanism of C–N bonding. Initially, we used in situ electrochemical infrared spectroscopy to detect the reaction intermediates of the InN3 catalyst for electrocatalytic reduction of CO2 and dimethylamine to synthesize DMF. By applying voltage, we found that in situ electrochemical infrared spectroscopy detected intermediates corresponding to C–O, C–N, and COOH bonds at wavelengths of 1173 cm−1, 1449 cm−1, and 1664 cm−1,26 respectively, indicating that these reaction intermediates are likely to be key reaction intermediates for the synthesis of DMF (Fig. 4b and e). However, the InN4 single-atom catalyst did not have a peak for C–N bonds during the electrocatalytic reduction of CO2 and dimethylamine, and only exhibited a peak for the *COOH intermediate at a wavelength of 1654 cm−1 (Fig. 4c and f).26 It suggests that InN4 is not conducive to C–N bonding. By comparing the in situ infrared spectra of different nitrogen-coordinated indium single-atom catalysts for the catalytic reduction of CO2 and dimethylamine C–N bonding, it was found that the low nitrogen-coordinated indium single-atom catalyst InN3 is more conducive to C–N bonding. The reason may be that the low nitrogen-coordinated indium single-atom catalyst of InN3 is more conducive to the adsorption of nitrogen-containing reactant dimethylamine and thus more conducive to C–N coupling with CO2 reduction intermediates to form DMF, while InN4 is not conducive to the adsorption of dimethylamine and only reduction of carbon dioxide, the reduction intermediates of carbon dioxide are not conducive to C–N coupling with dimethylamine. Our experimental result on electrocatalytic synthesis of DMF is consistent with the conclusion obtained from in situ infrared spectroscopy. Compared with InN4, InN3 electrocatalytically reduces dimethylamine and CO2 to synthesize N,N-dimethylformamide in the highest yield.
We further investigated the reaction mechanism of efficient C–N bonding between CO2 and dimethylamine through in situ Raman spectroscopy using the indium single-atom catalyst of InN3 for the electrochemical synthesis of DMF. The in situ Raman spectra of InN3 revealed peaks for the reactant dimethylamine at 890 cm−1 and 1018 cm−1, as shown in Fig. 4h, as well as peaks for the carbon material at 1357 cm−1 and 1602 cm−1. As the voltage increases, a C–N peak27 appears at 1064 cm−1 at –0.8 V (Fig. 4g and h). However, we detected a very small peak of C–N bond and did not observe any other Raman peaks of other reaction intermediates on InN4 (Fig. S10, ESI†), indicating that the InN4 catalyst is not easy to catalyze the C–N bonding of dimethylamine and CO2 to synthesize DMF. This is also consistent with our experimental results of in situ infrared spectroscopy and DFT calculations. At the same time, we also used in situ mass spectrometry detection as a supplementary method to investigate the mechanism of efficient C–N bonding electrochemical synthesis of DMF between CO2 and dimethylamine catalyzed by InN3. We found that the species COOH intermediate has a mass-to-charge ratio of m/z of 45, and did not find the species CO with an m/z value of 28 (Fig. 4i). Therefore, we believe that the *COOH intermediate is a key reaction intermediate for its C–N coupling. From in situ infrared spectroscopy and in situ online mass spectrometry, it is inferred that the reaction pathway for the C–N coupling of CO2 and dimethylamine to generate DMF is as follows: NH(CH3)2 + CO2 → *N(CH3)2 + *COOH → *COOHN(CH3)2 → *CON(CH3)2 → HCON(CH3)2.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ee05681g |
This journal is © The Royal Society of Chemistry 2025 |