The meso-substituent electronic effect of Fe porphyrins on the electrocatalytic CO2 reduction reaction

Hongyuan He , Zi-Yang Qiu , Zhiyuan Yin , Jiafan Kong , Jing-Shuang Dang *, Haitao Lei *, Wei Zhang and Rui Cao *
Key Laboratory of Applied Surface and Colloid Chemistry, Ministry of Education, School of Chemistry and Chemical Engineering, Shaanxi Normal University, Xi’an 710119, China. E-mail: dangjs@snnu.edu.cn; leiht2017@snnu.edu.cn; ruicao@snnu.edu.cn

Received 9th April 2024 , Accepted 8th May 2024

First published on 9th May 2024


Abstract

We report Fe porphyrins bearing different meso-substituents for the electrocatalytic CO2 reduction reaction (CO2RR). By replacing two and four meso-phenyl groups of Fe tetraphenylporphyrin (FeTPP) with strong electron-withdrawing pentafluorophenyl groups, we synthesized FeF10TPP and FeF20TPP, respectively. We showed that FeTPP and FeF10TPP are active and selective for CO2-to-CO conversion in dimethylformamide with the former being more active, but FeF20TPP catalyzes hydrogen evolution rather than the CO2RR under the same conditions. Experimental and theoretical studies revealed that with more electron-withdrawing meso-substituents, the Fe center becomes electron-deficient and it becomes difficult for it to bind a CO2 molecule in its formal Fe0 state. This work is significant to illustrate the electronic effects of catalysts on binding and activating CO2 molecules and provide fundamental knowledge for the design of new CO2RR catalysts.


Developing active and selective electrocatalysts for the carbon dioxide reduction reaction (CO2RR) has attracted great interest in the past few decades.1–4 First, the electrocatalytic CO2RR provides an appealing strategy to convert green electrical energy to chemical energy,5–14 which is stored in small molecules, such as CO, HCOOH, and CH4. On the other hand, by using the greenhouse gas CO2 as a C1 source, this process can produce value-added chemicals.15–19 Recent efforts have led to the development of a large variety of molecular complexes as active CO2RR electrocatalysts.20–28 Importantly, by studying these molecular catalysts, the correlation between the structure of the catalysts and their catalytic performance can be established.29–33 Among many studied structural factors, electronic effects were shown to play critical roles in determining catalytic activity and selectivity.34,35 However, understanding the electronic effects of catalysts on the CO2RR is still required from a fundamental aspect.

Fe porphyrins have been greatly studied as electrocatalysts for the CO2RR.36–44 Recent works from us and others showed that the activity of Fe porphyrins can be significantly boosted by modifying second-sphere environments of the Fe active site through the introduction of substituents with proton relay capability,45,46 hydrogen-bonding,47,48 and electrostatic features.49,50 In addition to these studies, we considered that Fe porphyrins are appealing models to study the electronic effects on the CO2RR. Porphyrin ligands are able to provide a rigid and stable Fe–N4 coordination environment. With this benefit, we can introduce substituents of different electronic effects to the ligand backbone to fine tune the electronic structure of Fe porphyrins, while maintaining the same Fe–N4 coordination environment. Note that keeping the same coordination environment of the metal center is essential to study the intrinsic electronic effect of the substituents on catalysis.

Herein, we report on the electronic effects of Fe porphyrins with different meso-substituents on the electrocatalytic CO2RR. We designed and synthesized FeTPP and its structural analogues FeF10TPP and FeF20TPP by replacing two and four meso-phenyl groups of FeTPP with pentafluorophenyl groups, respectively (Fig. 1). We showed that FeTPP and FeF10TPP are effective to catalyze the reduction of CO2 to CO in dimethylformamide (DMF) with an activity order of FeTPP > FeF10TPP, but FeF20TPP catalyzes the reduction of protons rather than CO2 under the same conditions. Importantly, results from both experimental and theoretical studies revealed that the introduction of strong electron-withdrawing pentafluorophenyl groups decreases the electron density of the Fe center, making the binding of a CO2 molecule at the corresponding Fe0 state more difficult. This work is therefore significant to provide insights into understanding the electronic effect of catalysts on the CO2RR, which will be valuable for the design and development of more efficient electrocatalysts.


image file: d4cc01630k-f1.tif
Fig. 1 Molecular structures of FeTPP, FeF10TPP, and FeF20TPP.

Fe porphyrins FeTPP, FeF10TPP, and FeF20TPP were synthesized according to the methods described in the ESI. High-resolution mass spectrometry analysis displayed an ion at a mass-to-charge ratio of 668.1646 for FeTPP (Fig. S4, ESI), 848.0709 for FeF10TPP (Fig. S5, ESI), and 1027.9770 for FeF20TPP (Fig. S6, ESI), confirming the identity and purity of these Fe porphyrins. The UV-vis absorption spectra of FeTPP, FeF10TPP, and FeF20TPP each displayed characteristic Soret and Q bands at almost identical positions (Fig. 2a), indicating the similar coordination environment of the Fe ion in the three complexes.


image file: d4cc01630k-f2.tif
Fig. 2 (a) UV-vis spectra of FeTPP, FeF10TPP, and FeF20TPP in DMF. (b) CVs of FeTPP, FeF10TPP, and FeF20TPP in DMF. Conditions: 0.1 M Bu4NPF6, 0.5 mM catalyst, 100 mV s−1 scan rate.

Next, the electrochemical features of these Fe porphyrins were tested in DMF. The cyclic voltammogram (CV) of FeTPP under argon displayed three reversible redox waves at −0.59, −1.51, and −2.13 V versus ferrocene, which could be assigned to the formal FeIII/II, FeII/I, and FeI/0 redox couple, respectively (Fig. 2b). Note that all potentials reported in this work are referenced to ferrocene. Similar results were observed for the other two Fe porphyrins, showing reversible redox waves at −0.49, −1.38, and −1.94 V for FeF10TPP and at −0.42, −1.24, and −1.81 V for FeF20TPP (Fig. 2b). These redox potentials are summarized in Table S1 (ESI). As compared to FeTPP, the reduction waves of FeF10TPP and FeF20TPP were notably shifted in the anodic direction. This result is consistent with the replacement of meso-phenyl groups in FeTPP with meso-pentafluorophenyl groups in FeF10TPP and FeF20TPP. In addition, in CV measurements, the reduction peak currents of all three Fe porphyrins increased linearly with the square root of the scan rates (Fig. S7–S9, ESI), indicating that these reduction processes are diffusion-controlled.

The electrocatalytic CO2RR performance of these Fe porphyrins was subsequently investigated in DMF at room temperature. As shown in Fig. 3a, the CV of FeTPP in CO2-saturated DMF with the addition of phenol as the proton source displayed a pronounced catalytic wave at the potential corresponding to the FeI/0 redox couple. The same measurement under argon in the presence of phenol showed no such catalytic wave. This result indicated that (1) this catalytic wave was due to the reduction of CO2 instead of protons and (2) the Fe0 form of FeTPP was the catalytically active species for the CO2RR. We also examined FeF10TPP and FeF20TPP for the electrocatalytic CO2RR in DMF. Our results showed that FeF10TPP was also active for the electrocatalytic CO2RR and the catalytic wave showed a shift of 220 mV to the anodic direction as compared to FeTPP (Fig. 3c). However, the catalytic current with FeF10TPP was much smaller than that with FeTPP. For FeF20TPP, it catalyzed the reduction of protons instead of the CO2RR (Fig. 3e) as its CVs in the presence of phenol under CO2 and under argon were similar to each other. These results suggested that strong electron-withdrawing meso-substituents of Fe porphyrins, such as pentafluorophenyl groups, are not favored for the CO2RR. Moreover, we determined the turnover frequency (TOF) of FeTPP and FeF10TPP for the electrocatalytic CO2RR using the method reported in the literature (Fig. S10, ESI).20,45 As shown in Table S2 (ESI), the TOF value is 124 s−1 for FeTPP and is 8.38 s−1 for FeF10TPP.


image file: d4cc01630k-f3.tif
Fig. 3 CVs of FeTPP (a), FeF10TPP (c), and FeF20TPP (e) in DMF under argon and CO2. CVs of FeTPP (b), FeF10TPP (d), and FeF20TPP (f) in acetonitrile under argon and CO2. Conditions: 0.1 M Bu4NPF6, 0.5 mM catalyst, 100 mV s−1 scan rate.

In addition to DMF, we also examined these Fe porphyrins for electrocatalytic CO2RR in acetonitrile. As shown in Fig. 3b and d, in the CVs of FeTPP and FeF10TPP, the FeI/0 wave under argon became a large catalytic wave under CO2 in the presence of phenol. Note that the catalytic CO2RR current obtained in acetonitrile is much higher than that in DMF, which is consistent with previous studies from us and others.51,52 Interestingly, for FeF20TPP, its CV with phenol showed little catalytic wave under argon but showed much larger catalytic current under CO2 (Fig. 3f), indicating that FeF20TPP became more selective for the CO2RR over hydrogen evolution in acetonitrile. Because illustrating the solvent effect on CO2RR selectivity is not the interest of the current work, we would like to not discuss the different electrocatalytic behaviors of FeF20TPP for the CO2RR in DMF and in acetonitrile herein.

Electrolysis under controlled potentials was then performed in DMF under CO2 at room temperature. For FeTPP, the electrolysis at −2.10 V showed stable currents at 0.48 mA cm−2 (Fig. S12, ESI). The UV-vis spectra of the DMF solution of FeTPP before and after electrolysis were almost identical (Fig. S12, ESI). The electrode after electrolysis displayed no electrocatalytic activity in a catalyst-free DMF solution under CO2 (Fig. S13, ESI). These results together confirmed that FeTPP is stable by functioning as a molecular electrocatalyst for the CO2RR. Gas chromatographic analysis confirmed the generation of CO during the electrolysis and gave a 98% faradaic efficiency for the CO2-to-CO conversion with FeTPP (Fig. S14, ESI). For FeF10TPP, it was also stable during the electrocatalytic CO2RR (Fig. S15, ESI) and gave a 94% faradaic efficiency for CO2-to-CO conversion (Fig. S16, ESI). However, for FeF20TPP, the faradaic efficiency for the CO2RR and for hydrogen evolution was 3% and 97%, respectively. These results are in an agreement with previous CV studies, showing that FeF20TPP has a poor selectivity for the CO2RR over hydrogen evolution.

To gain insights into the meso-substituent electronic effect on the CO2RR, we determined the CO2 binding feature of Fe porphyrins in DMF. By measuring the shift of the FeI/0 redox potential under CO2 and argon, we determined the CO2 binding constant (KCO2) using the method reported in the literature.53–55 No proton sources were added in the solution to prevent subsequent catalysis. With a high scan rate of 2.0 V s−1, the electron transfer and CO2 binding occurred quickly to achieve reversible kinetics (Fig. 4). Our results showed that the KCO2 value is 2.07 M−1 for FeTPP, 1.14 M−1 for FeF10TPP, and 0.26 M−1 for FeF20TPP (Table S2, ESI). We considered that with four strong electron-withdrawing meso-pentafluorophenyl substituents, the Fe0 state of FeF20TPP becomes highly electron-deficient to bind and activate a CO2 molecule.


image file: d4cc01630k-f4.tif
Fig. 4 CVs of FeTPP (a), FeF10TPP (b), and FeF20TPP (c) in DMF under argon and CO2. Conditions: 0.1 M Bu4NPF6 DMF, 0.5 mM catalyst, 2.0 V s−1 scan rate.

Density functional theory (DFT) calculations were then carried out to study the influence of meso-pentafluorophenyl substituents on the electronic structure and corresponding CO2 binding ability of these Fe porphyrins. First, we studied the electronic structure of the formal Fe0 state. As shown in Fig. S17 and Table S3 (ESI), the Fe0 species of all three Fe porphyrins exhibited triplet ground states with the spin localized on the central Fe atom, suggesting metal-centered reductions. Next, for CO2 binding at the Fe0 center, the calculated reaction free energy change was −5.6 kcal mol−1 for FeTPP, −4.3 kcal mol−1 for FeF10TPP, and −2.0 kcal mol−1 for FeF20TPP (Table S4, ESI). This result revealed that the CO2-binding ability of the Fe0 state of these Fe porphyrins has an order of FeTPP > FeF10TPP > FeF20TPP.

In addition, we illustrated in Fig. 5a the partial density of states (PDOS) distributions of the Fe0 center in the three Fe porphyrins, with the diagrams and the energy levels of the singly-occupied molecular orbital (SOMO). The results showed that the introduction of meso-pentafluorophenyl substituents leads to the reduction of the d orbital levels, which originated from their electron-withdrawing natures and can be verified by examining the electrostatic potential (ESP) maps (Fig. 5b). As a consequence, meso-pentafluorophenyl substituents cause an increased energy gap between the π* orbital of CO2 and the d orbitals of Fe (e.g. dz2 and dxz), leading to a reduced orbital interaction and thus a weakened intermolecular charge transfer efficiency between CO2 and the Fe0 center.


image file: d4cc01630k-f5.tif
Fig. 5 (a) PDOS distributions of Fe and the SOMO diagrams and energy levels of the Fe0 state of FeTPP, FeF10TPP, and FeF20TPP. (b) ESP maps of the Fe0 state of FeTPP, FeF10TPP, and FeF20TPP.

In conclusion, we report the meso-substituent electronic effect of Fe porphyrins on the electrocatalytic CO2RR. By introducing more meso-pentafluorophenyl substituents, Fe porphyrins became less active and selective for the CO2RR in DMF. We revealed that with strong electron-withdrawing meso-substituents, the central Fe atom of Fe porphyrins became electron-deficient, leading to decreased CO2 binding affinity of the corresponding Fe0 state. Therefore, although the introduction of electron-withdrawing substituents will make the catalytic wave shift in the anodic direction, it will also notably decrease the electron density of the metal ions, leading to decreased CO2 binding affinity.

We are grateful for support from the National Natural Science Foundation of China (22371177, 22171176 and 22325202), Key Research and Development Program of Shaanxi (2023-YBGY-296), Fok Ying-Tong Education Foundation for Outstanding Young Teachers in University, Fundamental Research Funds for the Central Universities (GK202306007 and GK202403004), and Research Funds of Shaanxi Normal University.

Conflicts of interest

There are no conflicts to declare.

Notes and references

  1. E. Boutin, L. Merakeb, B. Ma, B. Boudy, M. Wang, J. Bonin, E. Anxolabéhère-Mallart and M. Robert, Chem. Soc. Rev., 2020, 49, 5772–5809 RSC.
  2. Z. Liang, H.-Y. Wang, H. Zheng, W. Zhang and R. Cao, Chem. Soc. Rev., 2021, 50, 2540–2581 RSC.
  3. S. Lin, C. S. Diercks, Y.-B. Zhang, N. Kornienko, E. M. Nichols, Y. Zhao, A. R. Paris, D. Kim, P. Yang, O. M. Yaghi and C. J. Chang, Science, 2015, 349, 1208–1213 CrossRef CAS PubMed.
  4. D.-H. Nam, P. De Luna, A. Rosas-Hernández, A. Thevenon, F. Li, T. Agapie, J. C. Peters, O. Shekhah, M. Eddaoudi and E. H. Sargent, Nat. Mater., 2020, 19, 266–276 CrossRef CAS PubMed.
  5. M. R. Madsen, M. H. Rønne, M. Heuschen, D. Golo, M. S. G. Ahlquist, T. Skrydstrup, S. U. Pedersen and K. Daasbjerg, J. Am. Chem. Soc., 2021, 143, 20491–20500 CrossRef CAS PubMed.
  6. Y.-N. Gong, C.-Y. Cao, W.-J. Shi, J.-H. Zhang, J.-H. Deng, T.-B. Lu and D.-C. Zhong, Angew. Chem., Int. Ed., 2022, 61, e202215187 CrossRef CAS PubMed.
  7. G. Shi, Y. Xie, L. Du, X. Fu, X. Chen, W. Xie, T.-B. Lu, M. Yuan and M. Wang, Angew. Chem., Int. Ed., 2022, 61, e202203569 CrossRef CAS PubMed.
  8. Y. Cai, J. Fu, Y. Zhou, Y.-C. Chang, Q. Min, J.-J. Zhu, Y. Lin and W. Zhu, Nat. Commun., 2021, 12, 586 CrossRef CAS PubMed.
  9. S. Fernández, F. Franco, C. Casadevall, V. Martin-Diaconescu, J. M. Luis and J. Lloret-Fillol, J. Am. Chem. Soc., 2020, 142, 120–133 CrossRef PubMed.
  10. R. De, S. Gonglach, S. Paul, M. Haas, S. S. Sreejith, P. Gerschel, U.-P. Apfel, T. H. Vuong, J. Rabeah, S. Roy and W. Schöfberger, Angew. Chem., Int. Ed., 2020, 59, 10527–10534 CrossRef CAS PubMed.
  11. X. Rong, H.-J. Wang, X.-L. Lu, R. Si and T.-B. Lu, Angew. Chem., Int. Ed., 2020, 59, 1961–1965 CrossRef CAS PubMed.
  12. Y. Zhang, X.-Y. Zhang and W.-Y. Sun, ACS Catal., 2023, 13, 1545–1553 CrossRef CAS.
  13. A. A. Massie, C. Schremmer, I. Rüter, S. Dechert, I. Siewert and F. Meyer, ACS Catal., 2021, 11, 3257–3267 CrossRef CAS.
  14. M. E. Ahmed, A. Rana, R. Saha, S. Dey and A. Dey, Inorg. Chem., 2020, 59, 5292–5302 CrossRef CAS PubMed.
  15. C. Choi, S. Kwon, T. Cheng, M. Xu, P. Tieu, C. Lee, J. Cai, H. M. Lee, X. Pan, X. Duan, W. A. Goddard and Y. Huang, Nat. Catal., 2020, 3, 804–812 CrossRef CAS.
  16. F. Xu, B. Feng, Z. Shen, Y. Chen, L. Jiao, Y. Zhang, J. Tian, J. Zhang, X. Wang, L. Yang, Q. Wu and Z. Hu, J. Am. Chem. Soc., 2024, 146, 9365–9374 CrossRef CAS PubMed.
  17. J. Han, B. Tu, P. An, J. Zhang, Z. Yan, X. Zhang, C. Long, Y. Zhu, Y. Yuan, X. Qiu, Z. Yang, X. Huang, S. Yan and Z. Tang, Adv. Mater., 2024 DOI:10.1002/adma.202313926.
  18. T. Lu, T. Xu, S. Zhu, J. Li, J. Wang, H. Jin, X. Wang, J.-J. Lv, Z.-J. Wang and S. Wang, Adv. Mater., 2023, 35, 2310433 CrossRef CAS PubMed.
  19. B. Zhao, M. Sun, F. Chen, Y. Shi, Y. Yu, X. Li and B. Zhang, Angew. Chem., Int. Ed., 2021, 60, 4496–4500 CrossRef CAS PubMed.
  20. A. Sonea, N. R. Crudo and J. J. Warren, J. Am. Chem. Soc., 2024, 146, 3721–3731 CrossRef CAS PubMed.
  21. K. Teindl, B. O. Patrick and E. M. Nichols, J. Am. Chem. Soc., 2023, 145, 17176–17186 CrossRef CAS PubMed.
  22. W. Nie, D. E. Tarnopol and C. C. L. McCrory, J. Am. Chem. Soc., 2021, 143, 3764–3778 CrossRef CAS PubMed.
  23. M. H. Rønne, D. Cho, M. R. Madsen, J. B. Jakobsen, S. Eom, E. Éscoudé, H. C. D. Hammershøj, D. U. Nielsen, S. U. Pedersen, M.-H. Baik, T. Skrydstrup and K. Daasbjerg, J. Am. Chem. Soc., 2020, 142, 4265–4275 CrossRef PubMed.
  24. J. S. Derrick, M. Loipersberger, R. Chatterjee, D. A. Iovan, P. T. Smith, K. Chakarawet, J. Yano, J. R. Long, M. Head-Gordon and C. J. Chang, J. Am. Chem. Soc., 2020, 142, 20489–20501 CrossRef CAS PubMed.
  25. S. Dey, T. K. Todorova, M. Fontecave and V. Mougel, Angew. Chem., Int. Ed., 2020, 59, 15726–15733 CrossRef CAS PubMed.
  26. S. Sung, X. Li, L. M. Wolf, J. R. Meeder, N. S. Bhuvanesh, K. A. Grice, J. A. Panetier and M. Nippe, J. Am. Chem. Soc., 2019, 141, 6569–6582 CrossRef CAS PubMed.
  27. X. Liu, C. Liu, X. Song, X. Ding, H. Wang, B. Yu, H. Liu, B. Han, X. Li and J. Jiang, Chem. Sci., 2023, 14, 9086–9094 RSC.
  28. S. Gonell, E. A. Assaf, J. Lloret-Fillol and A. J. M. Miller, ACS Catal., 2021, 11, 15212–15222 CrossRef CAS.
  29. S. Dey, F. Masero, E. Brack, M. Fontecave and V. Mougel, Nature, 2022, 607, 499–506 CrossRef CAS PubMed.
  30. S. Amanullah, P. Gotico, M. Sircoglou, W. Leibl, M. J. Llansola-Portoles, T. Tibiletti, A. Quaranta, Z. Halime and A. Aukauloo, Angew. Chem., Int. Ed., 2024, 63, e202314439 CrossRef CAS PubMed.
  31. Y. Mao, M. Loipersberger, K. J. Kron, J. S. Derrick, C. J. Chang, S. M. Sharada and M. Head-Gordon, Chem. Sci., 2020, 12, 1398–1414 RSC.
  32. H. Guo, Z. Liang, K. Guo, H. Lei, Y. Wang, W. Zhang and R. Cao, Chin. J. Catal., 2022, 43, 3089–3094 CrossRef CAS.
  33. J. Han, N. Wang, X. Li, H. Lei, Y. Wang, H. Guo, X. Jin, Q. Zhang, X. Peng, X.-P. Zhang, W. Zhang, U.-P. Apfel and R. Cao, eScience, 2022, 2, 623–631 CrossRef.
  34. M. Tarrago, S. Ye and F. Neese, Chem. Sci., 2022, 13, 10029–10047 RSC.
  35. Y. Wang, X.-P. Zhang, H. Lei, K. Guo, G. Xu, L. Xie, X. Li, W. Zhang, U.-P. Apfel and R. Cao, CCS Chem., 2022, 4, 2959–2967 CrossRef CAS.
  36. J. S. Derrick, M. Loipersberger, S. K. Nistanaki, A. V. Rothweiler, M. Head-Gordon, E. M. Nichols and C. J. Chang, J. Am. Chem. Soc., 2022, 144, 11656–11663 CrossRef CAS PubMed.
  37. C. G. Margarit, N. G. Asimow, C. Costentin and D. G. Nocera, ACS Energy Lett., 2020, 5, 72–78 CrossRef CAS.
  38. A. Sonea, K. L. Branch and J. J. Warren, ACS Catal., 2023, 13, 3902–3912 CrossRef CAS.
  39. B. Mondal, P. Sen, A. Rana, D. Saha, P. Das and A. Dey, ACS Catal., 2019, 9, 3895–3899 CrossRef CAS.
  40. A. Khadhraoui, P. Gotico, W. Leibl, Z. Halime and A. Aukauloo, ChemSusChem, 2021, 14, 1308–1315 CrossRef CAS PubMed.
  41. R. Cao, ChemSusChem, 2022, 15, e202201788 CrossRef CAS PubMed.
  42. K. Guo, H. Lei, X. Li, Z. Zhang, Y. Wang, H. Guo, W. Zhang and R. Cao, Chin. J. Catal., 2021, 42, 1439–1444 CrossRef CAS.
  43. P. T. Smith, S. Weng and C. J. Chang, Inorg. Chem., 2020, 59, 9270–9278 CrossRef CAS PubMed.
  44. K. Torbensen, C. Han, B. Boudy, N. von Wolff, C. Bertail, W. Braun and M. Robert, Chem. – Eur. J., 2020, 26, 3034–3038 CrossRef CAS PubMed.
  45. C. Costentin, S. Drouet, M. Robert and J.-M. Savéant, Science, 2012, 338, 90–94 CrossRef CAS PubMed.
  46. S. Amanullah, P. Saha and A. Dey, J. Am. Chem. Soc., 2021, 143, 13579–13592 CrossRef CAS PubMed.
  47. M. R. Narouz, P. De La Torre, L. An and C. J. Chang, Angew. Chem., Int. Ed., 2022, 61, e202207666 CrossRef CAS PubMed.
  48. P. Gotico, B. Boitrel, R. Guillot, M. Sircoglou, A. Quaranta, Z. Halime, W. Leibl and A. Aukauloo, Angew. Chem., Int. Ed., 2019, 58, 4504–4509 CrossRef CAS PubMed.
  49. K. Guo, X. Li, H. Lei, H. Guo, X. Jin, X.-P. Zhang, W. Zhang, U.-P. Apfel and R. Cao, Angew. Chem., Int. Ed., 2022, 61, e202209602 CrossRef CAS PubMed.
  50. D. J. Martin and J. M. Mayer, J. Am. Chem. Soc., 2021, 143, 11423–11434 CrossRef CAS PubMed.
  51. K. Kosugi, M. Kondo and S. Masaoka, Angew. Chem., Int. Ed., 2021, 60, 22070–22074 CrossRef CAS PubMed.
  52. B. Zhao, H. Lei, N. Wang, G. Xu, W. Zhang and R. Cao, Chem. – Eur. J., 2020, 26, 4007–4012 CrossRef CAS PubMed.
  53. P. Gotico, L. Roupnel, R. Guillot, M. Sircoglou, W. Leibl, Z. Halime and A. Aukauloo, Angew. Chem., Int. Ed., 2020, 59, 22451–22455 CrossRef CAS PubMed.
  54. C. Cometto, L. Chen, P.-K. Lo, Z. Guo, K.-C. Lau, E. Anxolabéhère-Mallart, C. Fave, T.-C. Lau and M. Robert, ACS Catal., 2018, 8, 3411–3417 CrossRef CAS.
  55. E. M. Nichols, J. S. Derrick, S. K. Nistanaki, P. T. Smith and C. J. Chang, Chem. Sci., 2018, 9, 2952–2960 RSC.

Footnote

Electronic supplementary information (ESI) available: Experimental details; Fig. S1–S17; Tables S1–S4; Cartesian coordinates of calculated structures. See DOI: https://doi.org/10.1039/d4cc01630k

This journal is © The Royal Society of Chemistry 2024