A cobalt-tetraphenylporphyrin-based hypercrosslinked polymer for efficient CO2 photoreduction to CO

Saif Ullah a, Xunliang Hu a, Yaqin Zhang a, Irshad Hussain b, Xiaoyan Wang a, Hui Gao *a and Bien Tan *a
aKey Laboratory of Material Chemistry for Energy Conversion and Storage Ministry of Education, Hubei Key Laboratory of Material Chemistry and Service Failure, School of Chemistry and Chemical Engineering, Huazhong University of Science and Technology, Luoyu Road No. 1037, Wuhan, 430074, P. R. China. E-mail: bien.tan@mail.hust.edu.cn
bDepartment of Chemistry & Chemical Engineering, SBA School of Science & Engineering, Lahore University of Management Sciences (LUMS), DHA, Lahore Cantt, Lahore 54792, Pakistan

Received 9th October 2024 , Accepted 15th November 2024

First published on 18th November 2024


Abstract

One of the main impediments during the CO2 photocatalytic reduction process is controlling a particular product's selectivity while keeping a high conversion rate. The adsorption and affinity of the catalyst for CO2 are key factors for achieving highly efficient CO2 reduction. Herein, we report a novel photocatalytic material, cobalt anchored onto a tetraphenylporphyrin-based hypercrosslinked polymer (HCP-CoTPP), which was designed for the efficient photoreduction of CO2 to CO with an impressive evolution rate of 1449.9 μmol g−1 h−1 and nearly 100% selectivity in the presence of [Ru(Bpy)3]Cl2 as a photosensitizer and sacrificial agent under visible light irradiation.


Photocatalytic CO2 reduction is a compelling method for simultaneously utilizing solar energy and converting CO2 into valuable products.1,2 Substantial efforts have been dedicated to developing photocatalytic systems that can convert CO2 into a specific product efficiently.3 For efficient CO2 conversion, a system should ideally provide a platform with abundant catalytic centers and high porosity to support charge and mass transfer. This can be efficiently achieved by integrating catalytic sites into porous materials: the porosity promotes CO2 adsorption, while the catalytic sites facilitate its conversion.4,5

In recent years, various porous materials such as metal–organic frameworks (MOFs), covalent organic frameworks (COFs), and conjugated microporous polymers (CMPs) have demonstrated significant promise in constructing efficient photocatalytic CO2 reduction systems.6–8 Recently, Zheng et al.9 reported that photocatalytic CO2 reduction using GO-COF-366-Co achieved a CO selectivity of 96.1% and a yield of 52.2 mmol g−1. Rahimi et al.10 reported that compared to the non-donor 1,3,5-triethynylbenzene-based TEB-GFP CMP, photocatalytic CO2 reduction using a donor–acceptor based TPA-GFP CMP resulted in a 3-fold increase in CO evolution yield, reaching a maximum CO yield of 1666 μmol g−1 h−1 at 12 hours, and the selectivity improved significantly from 54% in TEB-GFP to 95% in TPA-GFP. However, these materials suffer from limitations such as high cost and difficult scalability.9,11–17 Hypercrosslinked polymers (HCPs), a class of porous organic polymers, are characterized by high surface area, tunable porosity, cost-effectiveness, ease of functionalization, and exceptional stability.18–21 Additionally, HCPs are ideal for incorporating catalytic sites, enabling the development of multiphase catalytic systems, making them highly promising for CO2 conversion applications.22–24 The effectiveness of CO2 conversion is closely linked to a material's CO2 capture ability.25 Materials with low adsorption capacities often suffer from reduced CO2 concentrations at the catalytic sites.26 However, most metal-based HCPs reported to date exhibit insufficient CO2 adsorption, leading to poor utilization of CO2.20,27,28 In this study, a tetraphenylporphyrin-based HCP (HCP-TPP) was synthesized via a Friedel–Crafts alkylation reaction. The heteroatoms in the porphyrin moiety exhibit CO2-philic properties, enhancing CO2 adsorption.29–32 Meanwhile, the porphyrin unit, recognized as a metal coordination site, facilitates the incorporation of metal ions, which act as catalytic centers while preventing metal ion agglomeration. Cobalt ions were subsequently introduced into the pores of HCP-TPP, enabling the photoreduction of CO2 to CO.33–35 The photocatalytic system exhibited outstanding photocatalytic performance, achieving a CO evolution rate of 1449.9 μmol g−1 h−1 with nearly 100% selectivity under visible light in the presence of Ru dye.12,36

The tetraphenylporphyrin-based HCP (HCP-TPP) was prepared via Friedel–Crafts alkylation reaction by the solvent kitting strategy under a nitrogen atmosphere with the temperature gradually increased from 0 °C to 80 °C (see ESI for experimental details). Cobalt ions were introduced into the HCP-TPP by a simple treatment of HCP-TPP with cobalt acetate, yielding HCP-CoTPP (Fig. 1).28 The chemical structures of HCP-TPP and HCP-CoTPP were analyzed by Fourier transform infrared (FT-IR) spectroscopy (Fig. 2a), and the bands at 1600 cm−1, 1500 cm−1, and 1450 cm−1 are assigned to the aromatic ring structure. The peaks in the area of 2920–2960 cm−1 are attributed to the characteristic aliphatic C–H stretching vibrations of the methylene group, implying the success of the Friedel–Crafts alkylation.37


image file: d4cc05331a-f1.tif
Fig. 1 Synthesis of a cobalt-tetraphenylporphyrin-based hyper crosslinked polymer.

image file: d4cc05331a-f2.tif
Fig. 2 (a) FT-IR spectra of HCP-TPP and HCP-CoTPP, (b) solid-state 13C NMR spectra of HCP-CoTPP, (c) XPS spectra for Co 2p of HCP-CoTPP, (d) nitrogen adsorption–desorption isotherms of HCPs at 77.3 K, (e) CO2 adsorption–desorption isotherm of HCPs at 273 K 1.13 bar, and (f) isosteric heat of adsorption of HCP-TPP and HCP-CoTPP.

Solid-state 13C cross polarization/magic angle spinning (CP/MAS) NMR spectroscopy was used to further confirm the chemical structure of HCP-CoTPP. The carbon atoms of the methylene groups are responsible for resonance peaks at about 52 ppm. Furthermore, the resonance peak assigned at 130 ppm due to the presence of aromatic carbon confirms the tetraphenylporphyrin structure (Fig. 2b).38

The chemical state of the Co was determined by X-ray photoelectron spectroscopy (XPS) (Fig. 2c). The binding energy peaks at 797.5 and 781.9 eV are assigned to Co2+ 2p1/2 and Co2+ 2p3/2, respectively, confirming that the valence state of the cobalt ion is +2 and the cobalt ions are successfully loaded into the tetraphenylporphyrin.39,40 The thermogravimetric analysis (TGA) of HCP-TPP and HCP-CoTPP under nitrogen demonstrated their good thermal stability (Fig. S1, ESI). The initial mass loss observed below 100 °C is primarily due to the presence of the entrapped gases and moisture in the high surface area HCPs. Scanning electron microscope (SEM) images show that the HCP-CoTPP has an irregular bulk morphology (Fig. S2, ESI). The energy-dispersive X-ray spectroscopy (EDX) elemental mapping demonstrated the homogeneous distribution of cobalt ions in HCP-CoTPP. The cobalt content, as measured by ICP-OES analysis, is 3.5 wt%.

The porous properties of the HCPs were analysed by nitrogen sorption analysis at 77.3 K. The detailed Brunauer–Emmett–Teller (BET) surface area, pore volumes, and micropore volumes of the materials are summarised in Table S1 (ESI). The specific surface areas of the HCP-TPP and HCP-CoTPP were 1239 and 861 m2 g−1, respectively. As illustrated in Fig. 2d, the adsorption isotherms of all samples exhibit a sharp nitrogen gas uptake at low relative pressure (P/P0 < 0.01), implying the presence of microporosity. The hysteresis loop at the medium pressure region (P/P0 = 0.4–0.7) reveals mesopores in the samples. The pore size distribution, calculated by nonlocal density functional theory (NLDFT), also confirms thepresence of micropores and mesopores in HCP-TPP and HCP-CoTPP (Fig. S3, ESI). CO2 uptake capability is an important parameter for highly efficient CO2 conversion. As shown in Fig. 2e, HCP-TPP exhibits the CO2 adsorption capability of 21.9 wt% (1.13 bar, 273 K) and 13.5 wt% (1.13 bar, 298 K). The incorporation of Co2+ ions into HCP-TPP led to a reduction in CO2 adsorption capacity, primarily due to a decrease in specific surface area and micropore volume. The CO2 uptake of HCP-CoTPP decreased to 13.1 wt% at 1.13 bar/273 K and 8.2 wt% at 1.13 bar/298 K (Fig. S4 and Table S2, ESI). To gain deeper insight into the adsorption interactions between tetraphenylporphyrin-based HCPs and CO2 molecules, the isosteric heat of adsorption (Qst) was calculated from CO2 adsorption isotherms at 273 K and 298 K using the Clausius–Clapeyron equation (Fig. 2f). The onset Qst values of HCP-TPP and HCP-CoTPP were 31.2 and 29.2 kJ mol−1, respectively, which is higher than that of the most of the porous organic polymers, such as TpBpy-COF (22 kJ mol−1), Por-sp2c-COF (25 kJ mol−1) and PCP-Cl (27 kJ mol−1).41–43 The high Qst values indicate strong interactions and high affinity between CO2 and the tetraphenylporphyrin-based HCPs, which promotes subsequent conversion reactions.

Given the catalytic properties of Co–ligand complexes for CO2 reduction and the high CO2 affinity of HCP-CoTPP, we evaluated the photocatalytic CO2 conversion activity of HCP-CoTPP. This was performed in the presence of [Ru(Bpy)3]Cl2 (Bpy = 2,2′-bipyridine) as a visible-light photosensitizer and triethanolamine (TEOA) as a hole-sacrificial reagent, in an acetonitrile/aqueous solution under visible light irradiation (λ ≥ 420 nm, 300 W Xe light source). HCP-CoTPP exhibited the CO evolution rate of 103.6 μmol g−1 h−1, while with [Ru(Bpy)3]Cl2, the CO evolution rate was 1449.9 μmol g−1 h−1 (Fig. 3a). When DMF was used as the solvent, the photocatalytic efficiency was improved to 2579.9 μmol g−1 h−1 (Fig. S5, ESI). The H2 evolution is negligible, which gives nearly 100% selectivity. The photocatalytic performance of HCP-CoTPP is comparable with other previously reported porous organic photocatalysts (Table S3, ESI).


image file: d4cc05331a-f3.tif
Fig. 3 (a) The photocatalytic CO2 conversion rate under different conditions using CoTPP as a catalyst and visible light irradiation (λ ≥ 420 nm). Condition 1: under N2, conditions 2 and 3: under CO2, (b) mass spectrum of 13CO produced using HCP-CoTPP as the catalyst in the photocatalytic reduction of 13CO2, (c) apparent quantum yield (AQY) of CO2 photocatalytic reduction for HCP-CoTPP at different wavelengths, (d) CO evolution by HCP-CoTPP (10 mg) with the presence of [Ru(Bpy)3]Cl2 = (10 mg) within 6 hours.

Control experiments were carried out by replacing the CO2 with N2, in which CO was not detected in the photocatalytic reaction under similar conditions either with or without photosensitizer (Fig. 3a). To confirm the origin of CO, an isotope labeling study was conducted using 13CO2 as a carbon source. The peak at m/z = 29 indicates that the generated 13CO is from 13CO2, demonstrating that the 13CO evolution from 13CO2 not from the decomposition of any other organic compounds in the photocatalytic system (Fig. 3b). Apparent quantum yield (AQY) was measured at different wavelengths, showing 0.10% at 420 nm, 0.33% at 450 nm, 0.35% at 500 nm, and 0.08% at 550 nm (Fig. 3c). Furthermore, CO production continued for up to 6 hours, indicating that the catalyst remains stable during the photocatalytic reaction system (Fig. 3d). HCP-CoTPP demonstrates excellent durability, achieving a high yield of CO evolution over 24 hours under visible light irradiation (Fig. S6, ESI). From the NMR spectra of the reaction solution (Fig. S7 and S8, ESI), no carbon-containing products were detected in the reaction solution.

A possible reaction mechanism for the photocatalytic reduction of CO2 over HCP-CoTPP is proposed (Fig. S9, ESI). The process involves the photoexcitation of [Ru(Bpy)3]Cl2 under light irradiation, generating excited electrons and holes. The excited electrons are transferred to the HCP-CoTPP and coupled with CO2 to produce CO. The holes are reductively quenched by TEOA.39,44,45

In conclusion, the cobalt-tetraphenylporphyrin-based HCP was prepared by a simple two-step strategy. HCP-CoTPP exhibited high CO2 uptake (13.1 wt% at 273 K, 1.13 bar), strong interaction (Qst: 29.2 kJ mol−1) and high affinity with CO2, and had abundant catalytic sites. The photocatalytic CO2 reduction by HCP-CoTPP is efficient in the presence of photosensitizer and a hole-sacrificial reagent. This photocatalytic system demonstrated a CO evolution rate of 1449.9 μmol g−1 h−1, and selectivity nearly 100% under visible light. This study demonstrates that the good CO2 uptake and high affinity to CO2 improve its photocatalytic conversion efficiency, providing new insight for developing highly efficient CO2 photocatalytic systems.

The authors acknowledge the assistance of the Analysis and Testing Center team at Huazhong University of Science and Technology for their support in material characterization. This work was financially supported by the National Natural Science Foundation of China (Grant No. 22161142005, 22475076) and National Key R&D Program of China (2023YFB4104000).

Data availability

The data supporting this article have been included as part of the ESI.

Conflicts of interest

There are no conflicts to declare.

Notes and references

  1. A. Rawat, A. Kumar, A. Pal, D. Arukha, P. Joshi and P. Mohanty, ACS Appl. Nano Mater., 2024, 7, 3761–3773 CrossRef CAS .
  2. P. Joshi, S. Mehta, N. Singh, S. Dalakoti, S. Divekar, S. Dasgupta, M. Srivastava and O. P. Khatri, J. Environ. Chem. Eng., 2023, 11, 110291 CrossRef CAS .
  3. S. Wang, X. Hai, X. Ding, S. Jin, Y. Xiang, P. Wang, B. Jiang, F. Ichihara, M. Oshikiri and X. Meng, Nat. Commun., 2020, 11, 1149 CrossRef PubMed .
  4. A. Rawat, R. Muhammad, V. Chandra Srivastava and P. Mohanty, Macromolecules, 2023, 56, 1236–1242 CrossRef CAS .
  5. J. Li, K. Ma, Y. He, S. Ren, C. Li, X.-B. Chen, Z. Shi and S. Feng, Catal. Sci. Technol., 2021, 11, 7300–7306 RSC .
  6. M. Zhang, P. Huang, J. P. Liao, M. Y. Yang, S. B. Zhang, Y. F. Liu, M. Lu, S. L. Li, Y. P. Cai and Y. Q. Lan, Angew. Chem., Int. Ed., 2023, 62, e202311999 CrossRef CAS PubMed .
  7. K. Seob Song, P. W. Fritz, D. F. Abbott, L. Nga Poon, C. M. Caridade, F. Gándara, V. Mougel and A. Coskun, Angew. Chem., Int. Ed., 2023, 62, e202309775 CrossRef CAS PubMed .
  8. Z. Ding, X. Tang, D. Zhao, S. Yan, L. Li, P. Li, W. Tang, S.-Y. Zhang and Y.-J. Zeng, Inorg. Chem., 2024, 63, 14193–14199 CrossRef CAS PubMed .
  9. Y. N. Gong, J. H. Mei, W. J. Shi, J. W. Liu, D. C. Zhong and T. B. Lu, Angew. Chem., Int. Ed., 2024, 63, e202318735 CrossRef CAS PubMed .
  10. F. A. Rahimi, A. Singh, R. Jena, A. Dey and T. K. Maji, ACS Appl. Mater. Interfaces, 2024, 16, 43171–43179 CrossRef CAS PubMed .
  11. J. Yang, Z. Chen, L. Zhang and Q. Zhang, ACS Nano, 2024, 18, 21804–21835 CrossRef CAS PubMed .
  12. D. H. Streater, E. R. Kennehan, D. Wang, C. Fiankor, L. Chen, C. Yang, B. Li, D. Liu, F. Ibrahim and I. Hermans, J. Am. Chem. Soc., 2024, 146, 4489–4499 CrossRef CAS PubMed .
  13. K. Sun, Y. Huang, Q. Wang, W. Zhao, X. Zheng, J. Jiang and H.-L. Jiang, J. Am. Chem. Soc., 2024, 146, 3241–3249 CrossRef CAS PubMed .
  14. Y. Zheng, X. Shen, M. Lin, M. Zhu, B. Yang, J. Yan, Z. Zhuang and Y. Yu, Small, 2024, 20, 2306836 CrossRef CAS PubMed .
  15. N. Li, G.-Q. Lai, L.-H. Chung, F. Yu, J. He and Y.-Q. Lan, CCS Chem., 2024, 6, 1211–1221 CrossRef CAS .
  16. P. Chakrabortty, S. Ghosh, R. V. Singh, M. R. Pai, K. Yusuf and S. M. Islam, ACS Sustainable Chem. Eng., 2024, 12, 10978–10992 CrossRef CAS .
  17. R. K. Saravanan, S. Karmakar, F. A. Rahimi, A. Dey, R. Jena, D. Maity and T. K. Maji, Chem. Mater., 2024, 36, 6410–6420 CrossRef CAS .
  18. G. E. Schukraft, R. T. Woodward, S. Kumar, M. Sachs, S. Eslava and C. Petit, ChemSusChem, 2021, 14, 1720–1727 CrossRef CAS PubMed .
  19. S. Wang, M. Xu, T. Peng, C. Zhang, T. Li, I. Hussain, J. Wang and B. Tan, Nat. Commun., 2019, 10, 676 CrossRef CAS PubMed .
  20. Z. Zhan, H. Wang, Q. Huang, S. Li, X. Yi, Q. Tang, J. Wang and B. Tan, Small, 2022, 18, 2105083 CrossRef CAS PubMed .
  21. F. Meng, J. Wang, M. Chen, Z. Wang, G. Bai and X. Lan, ACS Catal., 2023, 13, 12142–12152 CrossRef CAS .
  22. Y. Ma, X. Yi, S. Wang, T. Li, B. Tan, C. Chen, T. Majima, E. R. Waclawik, H. Zhu and J. Wang, Nat. Commun., 2022, 13, 1400 CrossRef CAS PubMed .
  23. T.-Y. Yang, Y. Zhang, G.-L. Zhang, J.-J. Zhang and Y.-H. Zhang, Catal. Lett., 2023, 153, 408–418 CrossRef CAS .
  24. H. Lv, R. Sa, P. Li, D. Yuan, X. Wang and R. Wang, Sci. China: Chem., 2020, 63, 1289–1294 CrossRef CAS .
  25. H. Zhou, C. Rayer, A. R. Antonangelo, N. Hawkins and M. Carta, ACS Appl. Mater. Interfaces, 2022, 14, 20997–21006 CrossRef CAS PubMed .
  26. X. Chi, Z. A. Lan, Q. Chen, X. Zhang, X. Chen, G. Zhang and X. Wang, Angew. Chem., Int. Ed., 2023, 62, e202303785 CrossRef CAS PubMed .
  27. H. Ouyang, M. Peng, K. Song, S. Wang, H. Gao and B. Tan, Chem. Synth., 2024, 4, 50 Search PubMed .
  28. H. Ouyang, K. Song, J. Du, Z. Zhan and B. Tan, Chem. Eng. J., 2022, 431, 134326 CrossRef CAS .
  29. J. Li, Y. Hou, C.-X. Cui, X. Zhang, J.-C. Wang, A. Wang, Z. Chen, M. Li and T. Lou, Dalton Trans., 2022, 51, 15022–15030 RSC .
  30. S.-J. Li, Y.-B. Chang, M. Li, Y.-X. Feng and W. Zhang, RSC Adv., 2020, 10, 17951–17954 RSC .
  31. X. Zhang, M. Cibian, A. Call, K. Yamauchi and K. Sakai, ACS Catal., 2019, 9, 11263–11273 CrossRef CAS .
  32. R. Luo, M. Chen, F. Zhou, J. Zhan, Q. Deng, Y. Yu, Y. Zhang, W. Xu and Y. Fang, J. Mater. Chem. A, 2021, 9, 25731–25749 RSC .
  33. Q. Chen, Y. Zhang, E. You, Q. Jiang, X. Chen, Y. Wang, Z. Song, K. Chang, Z. Xie and Q. Kuang, Small, 2022, 18, 2204924 CrossRef CAS PubMed .
  34. Y. Wu, L. Yan, Y. Yu and C. Jing, Catal. Sci. Technol., 2021, 11, 6103–6111 RSC .
  35. W. Y. Gao, H. T. Ngo, Z. Niu, W. Zhang, Y. Pan, Z. Yang, V. R. Bhethanabotla, B. Joseph, B. Aguila and S. Ma, ChemSusChem, 2020, 13, 6273–6277 CrossRef CAS PubMed .
  36. N. Das, R. Paul, S. Biswas, R. Das, R. Chatterjee, A. Bhaumik, S. C. Peter, B. M. Wong and J. Mondal, ACS Sustainable Chem. Eng., 2023, 11, 2066–2078 CrossRef CAS .
  37. H. Ouyang, K. Song, I. Hussain and B. Tan, ACS Appl. Polym. Mater., 2023, 5, 3574–3584 CrossRef CAS .
  38. W. Xu, Z. Zhang, Y. Wu, K. Chen and R. Luo, Chem. Commun., 2024, 60, 1599–1602 RSC .
  39. X. Hu, L. Zheng, S. Wang, X. Wang and B. Tan, Chem. Commun., 2022, 58, 8121–8124 RSC .
  40. S. Wang, K. Song, C. Zhang, Y. Shu, T. Li and B. Tan, J. Mater. Chem. A, 2017, 5, 1509–1515 RSC .
  41. W. Zhong, R. Sa, L. Li, Y. He, L. Li, J. Bi, Z. Zhuang, Y. Yu and Z. Zou, J. Am. Chem. Soc., 2019, 141, 7615–7621 CrossRef CAS PubMed .
  42. O. Buyukcakir, S. H. Je, D. S. Choi, S. N. Talapaneni, Y. Seo, Y. Jung, K. Polychronopoulou and A. Coskun, Chem. Commun., 2016, 52, 934–937 RSC .
  43. X. Kang, B. Wang, K. Hu, K. Lyu, X. Han, B. F. Spencer, M. D. Frogley, F. Tuna, E. J. McInnes and R. A. Dryfe, J. Am. Chem. Soc., 2020, 142, 17384–17392 CrossRef CAS PubMed .
  44. R. Paul, R. Das, N. Das, S. Chakraborty, C. W. Pao, Q. Thang Trinh, G. K. Kalhara Gunasooriya, J. Mondal and S. C. Peter, Angew. Chem., Int. Ed., 2023, 62, e202311304 CrossRef CAS PubMed .
  45. W. Zhao, D. Zhai, C. Liu, D. Zheng, H. Wu, L. Sun, Z. Li, T. Yu, W. Zhou and X. Fang, Appl. Catal., B, 2022, 300, 120719 CrossRef CAS .

Footnote

Electronic supplementary information (ESI) available: Experimental details, and additional characterization figures. See DOI: https://doi.org/10.1039/d4cc05331a

This journal is © The Royal Society of Chemistry 2024
Click here to see how this site uses Cookies. View our privacy policy here.