Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Light intensity–directed selective CO2 photoreduction using iron(0)–zirconium dioxide photocatalyst

Tomoki Oyumi, Ikki Abe, Masahito Sasaki and Yasuo Izumi*
Department of Chemistry, Graduate School of Science, Chiba University, Yayoi 1-33, Inage-ku, Chiba 263-8522, Japan. E-mail: yizumi@faculty.chiba-u.jp

Received 3rd March 2025 , Accepted 9th May 2025

First published on 13th May 2025


Abstract

Selective photoreduction of CO2 to CO, CH4, and C2,3 paraffins was directed by increasing ultraviolet-visible light intensity over an Fe–ZrO2 photocatalyst. Fe0 nanoparticles sequentially reduced COH—transferred from the ZrO2 surface—into CHx species and hydrocarbons, facilitated by light-induced heating to ∼452 K.


Photocatalytic CO2 reduction establishes a novel C-neutral cycle and is considered a potential environmental solution for a sustainable society.1,2 However, its near-term implementation is hindered by economic challenges, primarily due to the costs of photocatalysts and reactor design.3,4 Among first-row transition metals, Fe is the most abundant and cost-effective. Consequently, Fe-based photocatalysts for CO2 photoreduction have been widely studied. However, nearly all reported systems employ Fe in the form of metal ions within metal–organic frameworks, covalent organic frameworks, porphyrins, or as Fe2O3, functioning primarily as redox mediators (Table S1, ESI).5

In this study, a Fe0 surface was evaluated as a CO2 photoreduction catalyst for C1–3 hydrocarbons (HCs) in combination with semiconductor ZrO2. An Fe3O4–ZrO2 composite was synthesized via a liquid-phase reduction method, using Fe(NO3)3·9H2O as the precursor and reducing it at 973 K under H2 to obtain the Fe0 (7.5 wt%)–ZrO2-973R photocatalyst. The valence state and coordination of Fe were monitored during synthesis using X-ray absorption near-edge structure (XANES) and extended X-ray absorption fine structure (EXAFS).

This approach facilitated the formation of bifunctional sites—O vacancies (image file: d5cc01147g-t1.tif) on the ZrO2 surface3,6–9 and Fe0 nanoparticles—enabling the selective photogeneration of CO and C1–3 HCs from CO2. At a ultraviolet-visible (UV-Vis) light intensity of 110 mW cm−2, using 13CO2 and H2 gases with the Fe0 (7.5 wt %)–ZrO2-973R photocatalyst, both 13CO and 13CH4 were gradually generated over the first 5 h of photoreaction (Table 1a). However, beyond this period, 13CO formation became predominant (>99 mol% selectivity; entry a′), as subsequent reaction steps from CO did not proceed (Scheme 1a). The steady photogeneration of CO was significantly faster than that observed with ZrO2 alone (Table S2a, ESI),6–8 confirming the active role of the Fe0 surface, but the selectivity change after 5 h seemed due to strong CO2 adsorption on it in CO2 photoreduction (see below).

Table 1 CO2 photoreduction outcomes using Fe (7.5 wt%)–ZrO2 prereduced at 973 K in the presence of either H2 or H2O
Entry Reactant Reductant Light intensity (mW) Stage of reaction test (h) Formation rate (μmol h−1 gcat−1)
13CO 13CH4 13C2H6 13C3H8 O2
a 13CO2 (2.3 kPa) H2 (21.7 kPa) 110 cm−2 0–5 3.7 3.8 0.052 0.018 <0.002
a′ 5–48 3.7 <0.002 <0.002 <0.002
b 322 cm−2 0–5 2.4 45 0.20
b′ 5–48 1.7 1.1 0.036
c 472 cm−2 0–5 69 170 2.1 0.30
c′ 5–48 18 20 2.3 0.56
c′′ (20–48) (2.7) (16) (2.6) (0.70)
d 1510 cm−2 (with water bath) 0–5 35 40 0.26 <0.002 <0.002
d′ 5–48 5.2 0.41 0.037
  CO CH4 C2H6 C3H8 O2
e CO2 (95 kPa) H2O (70 mL) 367 per cell 0–5 12 <0.002 <0.002 64
e′ 5–48 8.5 0.86 9.0



image file: d5cc01147g-s1.tif
Scheme 1 Temporary and steady CO2 photoreduction pathways directed by irradiated light intensity (110–472 mW cm−2) using the Fe0 (7.5 wt%)–ZrO2-973R photocatalyst in presence of H2 ((a)–(c) and (c′)) or H2O (a).

In stark contrast, increasing the irradiated light intensity to 322 mW cm−2 shifted the Fe0–ZrO2-973R photocatalyst from selective two-electron reduction to 13CO to predominant eight-electron reduction to 13CH4 production (>95 mol% selectivity; Table 1b). This shift occurred because the hydrogenation steps from CO proceeded rapidly under higher light intensity, facilitating the sequential reduction from CO2 to CO and ultimately to CH4. From a practical perspective, however, the 13CH4 light-induced synthesis using Fe0–ZrO2 requires further improvement, as photocatalytic activity declined after 5 h of photoreaction—more so than under 110 mW cm−2 irradiation (Table 1a′ and b′). As no CHx species were observed in the Fourier transform infrared (FTIR) spectrum under the conditions, this deactivation was attributed to Fe0 surface poisoning by strongly adsorbed CO2 (Scheme 1b).

To enhance the stability of the Fe0–ZrO2-973R photocatalyst, the irradiated light intensity was further increased to 472 mW cm−2. Under these conditions, the 13CH4 formation rate reached 170 μmol h−1 gcat−1 with >71 mol% selectivity, accompanied by 13CO formation at 69 μmol h−1 gcat−1 (>28 mol% selectivity) over 5 h of photoreaction (Fig. 1A and Table 1c). Over time, 13CH4 formation exhibited a turning point at ∼5 h of photoreaction (Fig. 1A), attributed to partial Fe0 site poisoning by intermediate species such as HCO2 and CHx. Beyond this period, the 13CH4 formation rate stabilized at 20 μmol h−1 gcat−1 with >49 mol% selectivity (Table 1c′ and Scheme 1c). Under the steady conditions, CO2 blocking on Fe0 sites should be less effective in contrast to the status under 110–322 mW cm−2 irradiation. This time-dependent product distribution confirmed a consecutive reaction pathway: CO2 reduction to CO, followed by CH4 formation, and subsequent conversion to C2,3 hydrocarbons. Between 20 and 48 h of photoreaction, the 13CH4 formation rate remained stable (>73 mol% selectivity), while 13C2H6 and 13C3H8 selectivity significantly increased to 15 mol% (total formation rate: 3.3 μmol h−1 gcat−1), effectively suppressing initial 13CO production (Table 1c′′ and Scheme 1c′). Notably, no C2H4 or C3H6 was detected. Furthermore, photocatalytic activity could be restored by a 1-h evacuation at 10−6 Pa under light, reactivating the consecutive reduction process from CO to CH4 (Fig. 1B).


image file: d5cc01147g-f1.tif
Fig. 1 (A) and (B) Time course of 13CO, 13CH4, 12CH4, 13C2H6, and 13C3H8 formation using Fe0 (7.5 wt%)–ZrO2-973R, 13CO2 (2.3 kPa), H2 (21.7 kPa), and UV-Vis light irradiation at 472 mW cm−2. (B) Comparison with the reactivation test (filled symbols) after 1 h of catalyst evacuation under UV-Vis light irradiation at the 5-h mark.

The other photocatalytic test was using CO2 (2.3 kPa), H2 (21.7 kPa), and the light irradiation at 1510 mW cm−2, but the quartz reactor was cooled with water bath (Table 1d–d′ and Chart S1, ESI). The 13CO and 13CH4 formation rates were 24–51% of corresponding rates without water cooling until 5 h of reaction (Table 1c and c′), while the decline after 5 h was more significant, especially for 13CH4 formation, strongly suggesting that charge separation in/on ZrO2 owing to light proceeded CO2 reduction while Fe nanoparticle surface at lower temperature in thermal equilibrium with ZrO2, reactor, and 2.5 L of water was deactivated for subsequent multiple hydrogenation earlier by adsorbed species, e.g. CO2. The increase of water temperature was minimal: from 295.2 to 295.5 K during the photocatalytic test for 48 h (Chart S1, ESI).

Steady photocatalytic CO2 reduction was also achieved using H2O as the reductant instead of H2 under UV-Vis light irradiation at 367 mW per cell. CO was continuously generated at a rate of 12–8.5 μmol h−1 gcat−1 (Table 1e and e′), proceeding more rapidly than the multiple hydrogenation steps required for CH4 formation over Fe0 (selectivity <9.2 mol%; Scheme 1a and Fig. S1, ESI). This was attributed to the predominant presence of H2O rather than H at the Fe0 surface, which favored CO generation over further hydrogenation to CH4.

The effects of UV-Vis light irradiation on the Fe0–ZrO2-973R photocatalyst were investigated. In the UV-Vis absorption spectra of the Fe3O4–ZrO2 sample (Fig. S2b, ESI), two absorption shoulders at 410 and 525 nm, attributed to Fe2+ and Fe3+ ions, were observed, whereas ZrO2 alone exhibited no visible light absorption (spectrum a). This spectral profile was consistent with the reported absorption spectrum of Fe3O4.10 Upon H2 treatment at 973 K, absorption extended across the entire visible region (spectrum c), and the spectrum remained similar after 48 h of photocatalytic 13CO2 reduction (spectrum d), confirming the formation and stability of Fe0 nanoparticles, which facilitated multiple hydrogenation steps from CO/COH species (Scheme 1c and c′). The Fe valence state assignments were corroborated by changes in the XANES spectrum, aligning with those of standard Fe0 and Fe3O4 (Fig. S3A(c) and (d), ESI).11 EXAFS Fourier transform analysis of Fe0–ZrO2-973R revealed interatomic pairs characteristic of Fe0 (Fig. S4, ESI), while X-ray diffraction showed an Fe (011) peak overlapping with peaks corresponding to monoclinic ZrO2 (Fig. S5, ESI), further supporting the formation of Fe0 active sites in the Fe0–ZrO2-973R photocatalyst.

The fluorescence emission spectra of Fe0–ZrO2-973R were measured under excitation at 200 nm (Fig. S6, ESI), a wavelength corresponding to an energy higher than the band gap of ZrO2 (Fig. S2a, ESI). The spectrum exhibited both an interband excitation–deexcitation peak centered at 367 nm (spectrum a) and intraband transition peaks associated with impurity levels, such as O vacancies and Hf (0.55 wt%) in/on ZrO2, appearing at 451, 468, 481, 491, 528, and 623 nm. These emissions were significantly suppressed upon the addition of Fe3O4 and further reduced with Fe0 nanoparticles (spectra b and c), indicating effective trapping of excited electrons at the conduction band (CB) of ZrO2.

Next, the reaction pathway from CO2 to CO and HCs is considered. The role of image file: d5cc01147g-t2.tif sites on the ZrO2 surface in CO2 adsorption and its subsequent photoreduction to OCOH and COH species was analyzed using density functional theory (DFT) calculations.9 The population of surface image file: d5cc01147g-t3.tif sites was evaluated to one per 44 nm2 based on 13CO2 exchange amount with preadsorbed 12CO2 on image file: d5cc01147g-t4.tif site (0.070 μmol; Table S3g and Fig. S7, ESI). The population of surface image file: d5cc01147g-t5.tif sites seems not vary much before and after photocatalytic test based on essentially negligibly-changing UV-visible and XRD data (Fig. S2c, d and S5a, b, ESI). Consequently, this study focuses on the critical steps enabling transient or sustained C1–3 photogeneration (Scheme 1b and c), specifically the conversion of COH and/or CO into C1–3 HCs over the Fe0 surface.

To identify the active sites responsible for these reaction steps, Fe K-edge EXAFS measurements were conducted on the Fe0–ZrO2-973R photocatalyst under CO2 and H2 exposure. Unexpectedly, ∼20% of the Fe0 sites reduced at 973 K were oxidized upon reaction with CO2 in the dark, as indicated by EXAFS analysis (Fig. 2A, 0 min). The spectral fit, obtained by convolving standard XANES spectra for Fe0 and FeO with an 8[thin space (1/6-em)]:[thin space (1/6-em)]2 mixing ratio (Fig. S3B(d), ESI), aligned with the EXAFS data. This oxidation is attributed to the formation of an M-shaped Fe2+–O–C([double bond, length as m-dash]Fe0)–O–Fe2+ species upon CO2 adsorption on Fe. DFT calculations further support the energetic stability of this species on the Fe0 (111) surface, with an adsorption energy of 0.92 eV and a Bader charge of +0.352 on Fe bonded to O, consistent with prior studies.12 At this stage, CO did not desorb spontaneously from Fe surface and was not detected.


image file: d5cc01147g-f2.tif
Fig. 2 Time-dependent Fourier transform of Fe K-edge EXAFS for Fe0 (7.5 wt%)–ZrO2-973R under CO2 (2.3 kPa), H2 (21.7 kPa), and UV-Vis light irradiation (322 mW cm−2). (A) During 75 min of illumination, (B) after the light was turned off, and (C) corresponding time-dependent evolution of the Debye–Waller factor for the Fe–Fe interatomic pair.

Upon UV-Vis light irradiation, the Fourier transform of EXAFS spectra (Fig. 2A, 0 min) showed that the Fe–O and Fe⋯Fe peaks corresponding to FeO were rapidly replaced by a metallic Fe–Fe peak at 0.21 nm (phase shift uncorrected; Fig. 2A, 12 min). This transformation indicates the reduction of Fe2+ in the M-shaped Fe2+–O–C([double bond, length as m-dash]Fe0)–O–Fe2+ species.

In the FTIR spectrum of Fe0–ZrO2-973R under 13CO2 and H2 (Fig. S8, ESI), shoulder peaks at 1584, 1396, and 1217 cm−1 corresponded to νas(OCO), νs(OCO), and δ(OH) bending vibrations, respectively, were attributed to monodentate and bridging bicarbonate species adsorbed on the surface.6–8 In contrast, the broader peaks centered at 1538 and 1261 cm−1 were tentatively assigned νas(OCO) and νs(OCO) stretching vibrations of bidentate carbonate species on ZrO2,13 as well as the M-shaped Fe–O–C([double bond, length as m-dash]Fe0)–O–Fe species proposed earlier, adsorbed at various sites on the Fe0 surface. This assignment aligns with DFT-calculated O–C–O bond angles of 122°. Upon UV-Vis light irradiation (265 mW cm−2), the peak associated with 13CO2 and H2 adsorption negatively shifted by 41 cm−1. This shift was attributed to the reduction of Fe2+ to Fe0 via electron transfer from ZrO2 CB to Fe, followed by electron injection into the σ* orbitals of C–O bonds, facilitating bond weakening and activation under UV-Vis illumination, in consistent with the reduction from Fe2+ to Fe0 upon the light irradiation based on EXAFS (Fig. 2A).

The broad peak centered at 1497 cm−1 disappeared within 4 min under vacuum and UV-Vis light irradiation, likely due to CO2 desorption, suggesting that CO2 and carbonate species are not direct intermediates in C1–3 HC formation. In contrast, the conversion of adsorbed CO2 on the ZrO2 surface to COH species appears to be the rate-limiting step, as only methane (ν3 peak at 3010 cm−1) was detected alongside adsorbed CO2 and bicarbonate species.

To elucidate the energetic origins of the multiple hydrogenation steps in which COH and/or CO species migrate over the Fe0 surface to form HCs, the Fe–Fe coordination number was determined via EXAFS as 5.3, corresponding to a surface dispersion of 0.95.14 In the correlated Debye model, the Debye temperatures of bulk and surface Fe (vertical motion) are reported as 467 K15 and 225 K,16 respectively. Using multiple scattering calculation code FEFF17 and plane-wave EXAFS analysis code XDAP,18 the local Fe site temperature during CO2 photoreduction was experimentally monitored. Upon UV-Vis light irradiation, the Fe site temperature rapidly increased to ∼452 (±35) K as hot spot due to light absorption and remained nearly constant at thermal equilibrium with the supporting ZrO2 and the EXAFS cell both at ∼295 K (Fig. 2C, left).

When the light was turned off, the heat generated by light energy dissipated, and the Fe site temperature returned to 296 K (Fig. 2C, right). This observation confirms that the selective formation of C1–3 HCs in this study resulted from a two-step process: (i) CO2 reduction to COH/CO via charge separation in ZrO23,6–8 and (ii) subsequent multiple hydrogenation of CO/COH on the Fe0 surface, which was maintained at ∼452 K due to light absorption.

Thermal CO2 hydrogenation and Fischer–Tropsch synthesis of HC(s) using Fe-based catalysts typically require reaction temperatures above 548 K (Table S4, ESI) to achieve HC formation rates comparable to the Fe site temperature (∼452 K, Fig. 2C) and the observed rate of 0.24 mmol h−1 gcat−1 (Table 1c) in this study. Under CO2 photoreduction conditions, Fe sites remained exclusively in the Fe0 state (Fig. 2A and B), whereas thermal CO2 reduction commonly involves Fe oxides and carbides (Table S4, ESI). While the initial reduction of the first O atom in CO2 over Fe0 surfaces19 typically requires high temperatures (>548 K),20 this study demonstrates that the first reduction step was instead facilitated by charge separation at image file: d5cc01147g-t6.tif sites on the ZrO2 surface under light irradiation, enabling the reaction:

 
image file: d5cc01147g-t7.tif(1)
as previously reported.9 This step was particularly effective at lower light intensities (110–322 mW cm−2; Table 1a–b′) in the presence of neighboring Fe0 nanoparticles. The subsequent hydrogenation pathway, where COH migrates to Fe0 and undergoes further reduction to C1–3 HCs, is likely a common feature in both thermal (Table S4, ESI) and photo (∼452 K, Fig. 2C) reactions, regardless of whether H2 or H2O serves as the reductant.

The authors are grateful for the financial support from the Grant-in-Aid for Scientific Research B (24K01522, 20H02834, YI) from the Japan Society for the Promotion of Science. X-ray absorption experiments were performed with the approval of the Photon Factory Proposal Review Committee (2024G067, 2022G527, 2021G546). The authors would like to thank Enago (https://www.enago.jp) for the language review.

Data availability

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

Conflicts of interest

There are no conflicts to declare.

References

  1. Y. Izumi, Coord. Chem. Rev., 2013, 257, 171–186 CrossRef CAS.
  2. Y. Izumi, in Advances in CO2 Capture, Sequestration, and Conversion, ed. F. Jin, L.-N. He and Y. H. Hu, ACS Symposium Series, ACS Publications, 2015, ch. 1, vol. 1194, pp. 1–46 Search PubMed.
  3. T. Loumissi, R. Ishii, K. Hara, T. Oyumi, I. Abe, C. Li, H. Zhang, R. Hirayama, K. Niki, T. Itoi and Y. Izumi, Angew. Chem., Int. Ed., 2024, 63, e202412090 CrossRef CAS PubMed.
  4. J. Albero, Y. Peng and H. García, ACS Catal., 2020, 10, 5734–5749 CrossRef CAS.
  5. L. H. M. de Groot, A. Ilic, J. Schwarz and K. Wärnmark, J. Am. Chem. Soc., 2023, 145, 9369–9388 CrossRef CAS PubMed.
  6. H. Zhang, T. Itoi, T. Konishi and Y. Izumi, Angew. Chem., Int. Ed., 2021, 60, 9045–9054 CrossRef CAS PubMed.
  7. H. Zhang, T. Itoi, K. Niki, T. Konishi and Y. Izumi, Catal. Today, 2020, 356, 544–556 CrossRef CAS.
  8. H. Zhang, T. Itoi, T. Konishi and Y. Izumi, J. Am. Chem. Soc., 2019, 141, 6292–6301 CrossRef CAS PubMed.
  9. K. Hara, M. Nozaki, T. Hirayama, R. Ishii, K. Niki and Y. Izumi, J. Phys. Chem. C, 2023, 127, 1776–1788 CrossRef CAS.
  10. A. Bouafia, S. E. Laouini, A. Khelef, M. L. Tedjani and F. Guemari, J. Cluster Sci., 2021, 32, 1033–1041 CrossRef CAS.
  11. Provided by Dr. Y. Niwa, High Energy Accelerator Research Organization.
  12. C. R. Kwawu, R. Tia, E. Adei, N. Y. Dzade, C. R. A. Catlow and N. H. de Leeuw, Phys. Chem. Chem. Phys., 2017, 19, 19478–19486 RSC.
  13. S. E. Collins, M. A. Baltanas and A. L. Bonivardi, J. Catal., 2004, 226, 410–421 CrossRef CAS.
  14. B. J. Kip, F. B. M. Duivenvoorden, D. C. Koningsberger and R. Prins, J. Catal., 1987, 105, 26–38 CrossRef CAS.
  15. American Institute of Physics Handbook, ed. D. E. Gray, B. H. Billings, H. P. R. Frederikse, D. F. Bleil, R. B. Lindsay, R. K. Cook, J. B. Marion, H. M. Crosswhite and M. W. Zemansky, McGraw-Hill, New York, USA, 1972, 3rd edn, pp. 4–115 Search PubMed.
  16. D. P. Jackson, Surf. Sci., 1974, 43, 431–440 CrossRef CAS.
  17. L. Ankudinov, B. Ravel, J. J. Rehr and S. D. Condradson, Phys. Rev. B: Condens. Matter Mater. Phys., 1998, 58, 7565–7576 CrossRef.
  18. M. Vaarkamp, H. Linders and D. Koningsberger, XDAP Version 3.2.9, XAFS Services International, Woudenberg, The Netherlands, 2022 Search PubMed.
  19. L. Krauser, Q. W. Yang and E. V. Kondratenko, ChemCatChem, 2024, 16, e202301716 CrossRef.
  20. C. R. Kwawu, A. Aniagyei, R. Tia and E. Adei, Mater. Renew. Sustainable Energy, 2020, 9, 4 CrossRef.

Footnote

Electronic supplementary information (ESI) available: Lists of reported photo- and thermal catalysts, experimental details, photocatalytic time course, X-ray diffraction, and optical, X-ray absorption, and FTIR spectroscopy of photocatalyst and surface species. See DOI: https://doi.org/10.1039/d5cc01147g

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.