Next Article in Journal
Facile Recycling of Waste Biomass for Preparation of Hierarchical Porous Carbon with High-Performance Electromagnetic Wave Absorption
Previous Article in Journal
Assessment of Bitterness in Non-Charged Pharmaceuticals with a Taste Sensor: A Study on Substances with Xanthine Scaffold and Allopurinol
Previous Article in Special Issue
UV Laser-Induced Photodecomposition of Matrix-Isolated Salicylhydroxamic Acid: Identification of New Isocyanate Complexes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Methane Formation Induced via Face-to-Face Orientation of Cyclic Fe Porphyrin Dimer in Photocatalytic CO2 Reduction

by
Yusuke Kuramochi
1,2,*,†,
Masaya Hashimoto
1 and
Akiharu Satake
1,2,*
1
Department of Chemistry, Graduate School of Science, Tokyo University of Science, 1-3 Kagurazaka, Shinjuku-ku, Tokyo 162-8621, Japan
2
Department of Chemistry, Faculty of Science Division II, Tokyo University of Science, 1-3 Kagurazaka, Shinjuku-ku, Tokyo 162-8621, Japan
*
Authors to whom correspondence should be addressed.
Current address: Institute of Industrial Science, The University of Tokyo, 4-6-1 Komaba, Meguro-ku, Tokyo 153-8505, Japan.
Molecules 2024, 29(11), 2453; https://doi.org/10.3390/molecules29112453
Submission received: 22 April 2024 / Revised: 17 May 2024 / Accepted: 21 May 2024 / Published: 23 May 2024
(This article belongs to the Special Issue Recent Progress of Organic Photochemistry)

Abstract

:
Iron porphyrins are known to provide CH4 as an eight-electron reduction product of CO2 in a photochemical reaction. However, there are still some aspects of the reaction mechanism that remain unclear. In this study, we synthesized iron porphyrin dimers and carried out the photochemical CO2 reduction reactions in N,N-dimethylacetamide (DMA) containing a photosensitizer in the presence of 1,3-dimethyl-2-phenyl-2,3-dihydro-1H-benzo[d]imidazole (BIH) as an electron donor. We found that, despite a low catalytic turnover number, CH4 was produced only when these porphyrins were facing each other. The close proximity of the cyclic dimers, distinguishing them from a linear Fe porphyrin dimer and monomers, induced multi-electron CO2 reduction, emphasizing the unique role of their structural arrangement in CH4 formation.

Graphical Abstract

1. Introduction

The rise in atmospheric CO2 concentration has led to serious impacts on the environment, emphasizing the urgent requirement for CO2 reduction. Many attempts have been made to utilize light energy to reduce CO2 and convert it into energy-rich substances similar to photosynthesis. One significant challenge is the selective reduction of CO2 while suppressing thermodynamically favorable proton reduction, which generates hydrogen, and this has been addressed using metal complexes. For instance, in artificial Z-scheme systems using semiconductor photocatalysts, the incorporation of metal complex catalysts at reduction sites enables selective CO2 reduction under visible light irradiation. This process effectively extracts electrons from water, and simultaneously minimizes hydrogen generation, producing CO and formic acid [1,2]. Abundant earth elements have been used as the central metal; among them, Fe porphyrin complexes have been extensively studied owing to their high selectivity and activity [3,4]. The introduction of peripheral proton-donating groups (Fe-o-OH in Figure 1) considerably improved the catalytic activity of Fe porphyrins, resulting in selective CO production [5]. In addition, linking Fe porphyrins through an o-phenylene or a urea group enhanced CO production [6,7]. The Fe porphyrin substituted with p-trimethylammoniophenyl groups (Fe-p-TMA in Figure 1) also exhibited high CO production activity in the electrochemical CO2 reduction [8]. Interestingly, under photochemical CO2 reduction conditions using tris(2-phenylpyridine)iridium (Ir(ppy)3) as a photosensitizer and triethylamine (TEA) as a sacrificial electron donor in acetonitrile, Fe-o-OH, and Fe-p-TMA yielded CH4 along with CO [9]. The formation of CH4 was observed when using an organic photosensitizer, 3,7-(4-biphenyl)-1-naphthalene-10-phenoxazine (Phen2), instead of Ir(ppy)3 in N,N-dimethylformamide (DMF) solution containing Fe-p-TMA and sacrificial electron donors [10]. The reaction mechanism indicated the involvement of the Fe(II)–CO intermediate. However, although computational studies have proposed this reaction mechanism [11], the actual formation mechanism of CH4, an eight-electron reduction product of CO2, remains unclear.
We previously synthesized cyclic Zn porphyrin dimers connected via 2,2′-bipyridine (bpy) and isophthalamide (Zn2-CP2m) or terephthalamide linkers (Zn2-CP2p) [12]. Herein, we substituted the central Zn ions of porphyrins with Fe ions to yield two types of cyclic Fe porphyrin dimers with different distances between the porphyrin planes (Fe2-CP2m and Fe2-CP2p in Figure 1). The photocatalytic CO2 reduction, in the presence of Ir(ppy)3 [9] or Phen2 [10] as the photosensitizer and 1,3-dimethyl-2-phenyl-2,3-dihydro-1H-benzo[d]imidazole (BIH) [13] as the electron donor, surprisingly produced CH4 only when using the cyclic Fe porphyrin dimers (Fe2-CP2m and Fe2-CP2p). In contrast, a linear Fe porphyrin dimer (Fe2-P2) and monomeric Fe porphyrins did not produce CH4, indicating that the close proximity of two Fe porphyrins facing each other induced the multielectron reduction of CO2.

2. Results and Discussion

Fe2-CP2m and Fe2-CP2p were, respectively, prepared by demetallation of Zn2-CP2m and Zn2-CP2p [12] to obtain the corresponding free-base porphyrins, followed by the introduction of Fe ions. Fe2-P2 was prepared by the introduction of Fe ions into the corresponding free-base precursor [12]. In the UV–vis absorption spectra, the Soret band of Fe2-P2 red shifted compared with those of Fe2-CP2m, Fe2-CP2p, and Fe(III) tetraphenylporphyrin chloride (FeTPP(Cl)) (Figure 2, left). The red-shifted band is attributed to the head-to-tail excitonic coupling between the two transition dipoles of X in the anticonformation of Fe2-P2, indicating that a linear structure is stable in Fe2-P2, as observed for the corresponding Zn porphyrins (Figure 1) [12]. The cyclic voltammogram (CV) and differential pulse voltammogram (DPV) of Fe2-P2 in Ar-saturated DMF showed three reversible redox waves at −0.74, −1.55, and −2.23 V vs. Fc/Fc+ (−0.27, −1.08, and −1.76 V vs. SCE) [14], which corresponded, respectively, to Fe(III/II), Fe(II/I), and Fe(I/0) couples (Figure 2, right and Figure S12). The observation of only three redox waves indicates that each porphyrin is reduced independently and that the electronic interaction between the two porphyrins through the bpy linker is negligible. Meanwhile, although the CVs of Fe2-CP2m and Fe2-CP2p showed redox waves at the similar positions to those of Fe2-P2, the second redox wave of Fe2-CP2m split into two, indicating that the porphyrins in close proximity, arranged in a face-to-face configuration, exhibit electrical interaction with each other. Under a CO2 atmosphere, Fe2-P2, Fe2-CP2m, and Fe2-CP2p showed similar catalytic currents at the third waves of Fe(I/0) of the Fe porphyrin dimers (Figure 2, right). Lewis acids, known to enhance the activity of Fe porphyrin catalysts for CO2 reduction [15,16], were anticipated to exhibit differences in their interactions with the bpy moiety between cyclic and linear structures. However, no significant difference among three porphyrin dimers in the catalytic currents was observed for the electrochemical CO2 reduction even when using water and/or metal ion additives (Figures S13–S15 and Table S1).
We first attempted to perform photocatalytic CO2 reduction using Fe2-P2 in the presence of Ir(ppy)3 as a photosensitizer and TEA as an electron donor, but no reduction product was detected. The present Fe porphyrin dimers exhibited more negative reduction potentials (−2.17 to −2.23 V vs. Fc/Fc+ and −1.70 to −1.76 V vs. SCE for Fe(I/0)) compared with the reported Fe porphyrins, such as Fe-o-OH (−1.57 V vs. SCE for Fe(I/0)) and Fe-p-TMA (−1.47 V vs. SCE for Fe(I/0)) [17]. According to the energy diagram (Figure S16), the oxidation quenching process [18] of the excited Ir(ppy)3 by either Fe2-P2 or Fe2-CP2m is thermodynamically less favorable when using TEA. Consequently, we used BIH with a stronger reducing power as an electron donor instead of TEA. The energy diagram and results of a phosphorescence quenching experiment support that electron transfer from BIH to the excited Ir(ppy)3 (i.e., reductive quenching) [18], followed by a reduction in the Fe(I) porphyrins, can occur (Figures S17–S19). Here, BIH was used in much larger quantities (10 mM) than the catalyst (10 μM), and it was expected that the reaction could proceed, although it was slightly unfavorable thermodynamically. Under the reaction conditions for the photocatalytic CO2 reduction using BIH and Fe-o-OH, we observed the production of CO during the catalytic reaction (Table S2). However, the amount of BIH consumed was significantly greater than the production of CO (Figure S21). The NMR spectra after irradiation in the presence of CO2 revealed the formation of unidentified BIH decomposition products (Figure S22) rather than the formation of BI+, which is typically observed in reactions with the [Ru(bpy)3]2+ photosensitizer as a two-electron oxidation product of BIH [13]. A highly reactive BI radical, formed via the oxidation and deprotonation of BIH, is likely to react with CO2 to generate the unidentified products (Figure S23). This is hypothesized because Ir(ppy)3 would not accept the electron from the BI radical. The investigation of the effects of solvents and additives showed that N,N-dimethylacetamide (DMA) [19] promoted CO production and suppressed BIH consumption more than acetonitrile (entries 4−6 in Table S3 and Figure S24). Therefore, in the subsequent experiments, the photoreactions were performed using DMA.
Photocatalytic CO2 reductions using Fe2-P2 and Fe2-CP2m (10 μM) in DMA containing BIH (10 mM) and Ir(ppy)3 (0.2 mM) under 450 nm light were performed. The turnover numbers (TONs) of the reduction products against the Fe porphyrin dimers are shown in Figure 3. No detectable amounts of CO, H2, or CH4 were found in the absence of any one of the Fe porphyrin dimers, Ir(ppy)3, and light, whereas only H2 was detected under Ar instead of CO2. While Fe2-P2 and Fe2-CP2m produced CO, the amount of CO was smaller in Fe2-CP2m. Interestingly, a small amount of CH4 was formed in Fe2-CP2m (Figure 3b). Under the same conditions, CH4 was not detected in Fe2-P2 (Figure 3a), Fe-o-OH, or FeP-phen, which is a model monomer with a diimine ligand. Figure 3b shows CH4 production with an induction period and lower CO production than that in Figure 3a, indicating that the CH4 was formed via the reduction in CO, as reported previously [9,10].
Next, we investigated the photocatalytic CO2 reduction reactions in which the organic dye Phen2, rather than Ir(ppy)3, was used as the photosensitizer (Figure 4) [10]. During the irradiation of the solution in the absence of the catalyst (Figure S28), BIH decomposition was still observed without reduction products of CO2. However, reducing the light intensity to 5 mW suppressed BIH decomposition. When Fe2-P2 was used as the catalyst, a linear CO formation was observed for up to 4 h under the 5 mW light intensity (Figure S29). Figure 4 shows the TONs of the reduction products against the Fe porphyrin dimers in DMA containing BIH (10 mM) and Phen2 (1.0 mM) under 420 nm light (5 mW). As observed with Ir(ppy)3, CH4 was formed with induction periods in the cyclic structure (Fe2-CP2m and Fe2-CP2p), whereas no CH4 was detected in the linear structure (Fe2-P2). In addition, the CO productions in Figure 4b,c were smaller than those in Figure 4a. In the previous systems involving Fe-p-TMA as the catalyst and TEA as the electron donor, the addition of a proton source, such as trifluoroethanol (TFE), enhanced the formation of CO and CH4 [9,10]. However, using the cyclic porphyrin dimer, the CH4 production decreased with the addition of TFE and was completely suppressed by PhOH, while TFE and PhOH enhanced the CO production. The addition of Mg ions decreased the production of CO and CH4 (Figure S30).
In conventional catalytic reactions using Phen2 as the photosensitizer, the reaction typically involves an oxidative quenching process, where electrons are transferred from the excited Phen2 to the catalyst [18,20]. However, this system would proceed via a reductive quenching process involving electron transfer from the electron donor to the excited Phen2. The fluorescence quenching experiments of Phen2 by BIH demonstrate that the electron transfer from BIH to the excited singlet state of Phen was efficient (Figure S32). However, the quenching efficiency of the excited Phen2 (ηq), which was estimated from the Stern–Volmer plot [18], strongly depended on the concentration of BIH because of the shorter fluorescence lifetime ([BIH] = 10 mM, ηq = 7%; [BIH] = 100 mM, and ηq = 41%). Meanwhile, we observed that the TONs were less dependent on the BIH concentration (10−100 mM in Figure 5), suggesting that the electron transfer from BIH mainly occurred not via the excited singlet state but via the long-lived excited triplet state of Phen2 [21], which has a lifetime of 480 μs [22].
The photocatalytic CO2 reduction using Fe2-CP2p and 100 mM BIH produced CH4 with TON = 3.5 against the Fe porphyrin dimer after 35 h, exceeding the amount of the Fe porphyrin units (Figure S34). The capillary electrophoresis showed that the TON of formic acid reached 38 after irradiation for 18 h in the presence of Fe2-CP2p and BIH (100 mM, Figure S34). We conducted isotopic experiments under 12CO2 and 13CO2 atmospheres. In gas chromatography/mass spectrometry, 13CO (m/z = 29) and 13CH4 (m/z = 17) were detected under a 13CO2 atmosphere (Figures S35 and S36), confirming that the carbon source of CO and CH4 was CO2. The 1H and 13C NMR spectra were measured in a DMA-d9 solution during irradiation under 12CO2 and 13CO2 atmospheres (Figures S37–S39). The spectral changes during light irradiation showed that BIH was almost completely consumed after 21 h, indicating that the catalytic reaction stopped due to the disappearance of BIH. No reduction products, including methanol or formaldehyde, were observed, except for formic acid, which showed a doublet peak at 8.68 ppm with a coupling constant of J13C-H = 175 Hz and a singlet peak at 8.72 ppm in 1H NMR spectra under 12CO2 and 13CO2 atmospheres, respectively (Figure S38). A peak at 167 ppm was assigned to HC(O)-, which was correlated with the doublet proton peak at 8.68 ppm in the heteronuclear multiple bond connectivity (HMBC), observed in the 13C NMR spectrum under only a 13CO2 atmosphere (Figures S39 and S40) [19]. Formic acid was not observed in the absence of the Fe porphyrins, indicating that formic acid is produced via the CO2 reduction catalyzed by the Fe porphyrins, and it is not directly formed by the chemical reaction between BIH and CO2 [23,24]. Furthermore, a peak appeared at 222 ppm in the 13C NMR spectrum only under a 13CO2 atmosphere (indicated by an asterisk in Figure S39). The peak can be attributed to the Fe–13CO signal [25], indicating the formation of a carbonyl intermediate during irradiation, as observed in previous reports. In addition, only under a 13CO2 atmosphere an intense peak was observed at 172 ppm (Figure S39), which was correlated with the proton peaks at 7.4, 3.4, 2.7, 2.4, and 1.8 ppm in HMBC (Figure S40). Although no clear attribution could be established, it likely corresponded to the reaction products of BIH and CO2, an adduct of CO2 with the BI radical caused by the oxidation and deprotonation of BIH.

3. Materials and Methods

3.1. General Procedure

All chemicals and solvents were of commercial reagent quality and were used without further purification unless otherwise stated. Tris(2-phenylpyridine)iridium (Ir(ppy)3) was purchased from Sigma-Aldrich (St. Louis, MO, USA). Fe(III) 5,10,15,20-tetrakis(2,6-dihydroxyphenyl)porphyrin chloride (Fe-o-OH) [5], 3,7-(4-biphenyl)-1-naphthalene-10-phenoxazine (Phen2) [10], the cyclic Zn porphyrin dimers (Zn2-CP2m and Zn2-CP2p), the linear free-base porphyrin dimer (Fb2-P2) [12], 1,3-dimethyl-2-phenyl-2,3-dihydro-1H-benzo[d]imidazole (BIH) [13], and 5,10,15-tris(4-tert-butylphenyl)-20-(1,10-phenanthrolin-2-yl)-21H,23H-porphyrin (H2P-phen) [26] were prepared according to the literature. N,N-dimethylformamide (DMF) was dried over molecular sieves of size 4 Å. The reactions were monitored on silica gel 60F254 TLC plates (Merck, Darmstadt, Germany). The following silica gels utilized for the column chromatography were purchased from Kanto Chemical Co. Inc. (Tokyo, Japan): silica gels (Spherical, Neutral) 40–100 μm and (Flash) 40–50 μm. 1H, 13C NMR, and 1H-1H correlation spectroscopy (COSY), 1H−13C heteronuclear single quantum correlation (HSQC), and 1H−13C heteronuclear multiple bond correlation (HMBC) spectra were recorded using a JEOL JNM-ECZ-400 and a JEOL JNM-ECA-500 (JEOL, Tokyo, Japan). Chemical shifts were recorded in parts per million (ppm) relative to tetramethylsilane. MALDI–TOF mass spectra were collected on a JEOL JMS S-3000 with dithranol as a matrix with sodium iodide (NaI). UV-vis absorption spectra were collected using a square cell (path length = 1.0 cm) on a JASCO V-650 spectrometer (JASCO, Tokyo, Japan). The steady-state emission spectra were collected on an Hitachi F-4500 spectrometer and corrected for the response of the detector system. The fluorescence intensities were normalized at the absorbance of the excitation wavelength. Cyclic voltammogram (CV) and differential pulse voltammogram (DPV) were measured using an ALS-H/CHI Model 612E electrochemical analyzer (BAS, Tokyo, Japan) in a micro-cell equipped with a glassy carbon working electrode (ϕ 1.6 mm) and a Pt counter electrode. The micro-cell was connected via a Luggin capillary with a reference electrode of Ag/AgNO3 (10 mM in DMA). Tetrabutylammonium hexafluorophosphate (nBu4NPF6) recrystallized from ethyl acetate was used as a supporting electrolyte. Ferrocene was used as an external standard, and all potentials were referenced to the ferrocene/ferrocenium couple. The currents (i) were normalized by the peak currents (ip0) corresponding to the Fe(III)/Fe(II) wave in the absence of CO2 and additives. High-performance liquid chromatographies (HPLCs) were carried out using a JASCO PU-2089 and an MD-44010 system (JASCO, Tokyo, Japan) equipped with a TSKgel ODS-100S column (4.6 mm I.D. × 25 cm; Tosoh, Tokyo, Japan) using acetonitrile/H2O = 4/1 (v/v) as an eluent.

3.2. Synthesis of Fb2-CP2m

The TFA (0.50 mL) was slowly added to Zn2-CP2m (10 mg, 5.9 × 10−6 mol) in a 10 mL flask, and the mixture was stirred for 2 h. The solution was slowly poured into a saturated NaHCO3 aqueous solution in an ice bath. The organic layer was transferred to a PFA-coated funnel, and the aqueous layer was extracted with CHCl3 (10 mL × 3). The combined organic layer was washed with water and dried over anhydrous Na2SO4. The solvent was evaporated to dryness, resulting in a purple solid, Fb2-CP2m (8.3 mg, 90%). 1H NMR (500 MHz, pyridine-d5) δ/ppm = 10.85 (s, 2H, NH), 8.91 (s, 1H, Ph), 8.83 (m, 2H, Ph), 8.74 (m, 4H, β-pyrrole), 8.70 (m, 2H, Py), 8.58–8.47 (m, 12H, β-pyrrole, Ph, Py), 8.36 (m, 4H, β-pyrrole), 8.31 (m, 2H, Py), 8.11 (m, 2H, Ph), 7.74 (s, 2H, Ph), 7.67 (m, 2H, Ph), 7.55 (m, 1H, Ph), 7.27 (s, 4H, mesityl), 7.08 (s, 4H, mesityl), 2.48 (s, 12H, CH3), 1.64 (s, 12H, CH3), 1.54 (s, 12H, CH3), −2.84 (s, 4H, inner NH); MALDI-TOF mass: m/z [M + H]+ 1557.6808, calcd. for [C106H85N12O2]+ 1557.6913.

3.3. Synthesis of Fe2-CP2m

In a 10 mL flask were placed Fb2-CP2m (8.3 mg, 5.3 × 10−6 mol), CHCl3 (2.0 mL), and 2,6-lutidine (15 μL, 1.3 × 10−4 mol). Anhydrous FeCl2 (37 mg, 2.9 × 10−4 mol) dissolved in methanol (1.0 mL) was added to it, and the mixture was stirred under reflux for 18 h. The resulting solution was transferred to a perfluoroalkoxy alkane (PFA)-coated flask, and diluted with CHCl3 (ca. 10 mL). The organic solution was washed with 1 M HCl aqueous solution (×3) and water (×3) and then dried over anhydrous Na2SO4. The residue obtained by evaporation of the solvent was purified with a flush silica gel column (eluents: CHCl3, CHCl3:CH3OH = 50:1, and CHCl3:CH3OH = 10:1). The fraction eluted with CHCl3:CH3OH = 50:1 was collected and the solvent was evaporated. The solid was dissolved in CHCl3 (ca. 10 mL) was treated with 1 M HCl aqueous solution (×3) and water (×3), and passed through Phase Separator paper (Whatman, Maidstone, UK). The solvent was evaporated to dryness, giving the titled compound as a black solid (7.9 mg, 86%). MALDI-TOF mass: m/z [M − 2Cl + H]+ 1666.5287 (max), calcd. for [C106H81N12O2Fe2]+ 1666.5333; UV-vis absorption (CHCl3) λ max/nm (ε/M−1cm−1) = 373 (6.2 × 104, LMCT band), 415 (1.1 × 105, Soret band), and 509 (1.4 × 104, Q band).

3.4. Synthesis of Fb2-CP2p

The TFA (1.0 mL) was slowly added to Zn2-CP2p (8.7 mg, 5.2 × 10−6 mol) in a 10 mL flask, and the mixture was stirred for 2 h. The solution was slowly poured into a saturated NaHCO3 aqueous solution in an ice bath. The organic layer was transferred to a PFA-coated funnel and the aqueous layer was extracted with CHCl3 (10 mL × 3). The combined organic layer was washed with water and dried over anhydrous Na2SO4. The solvent was evaporated to dryness, giving a purple solid, Fb2-CP2p (7.1 mg, 89%). 1H NMR (400 MHz, CDCl3) δ/ppm = 8.67 (d, J = 8.2 Hz, 2H, Ph or Py), 8.63 (d, J = 7.8 Hz, 2H, Ph or Py), 8.52 (d, J = 4.7 Hz, 4H, β-pyrrole), 8.47 (t, J = 7.8 Hz, 2H, Ph or Py), 8.43–8.45 (2H, Ph or Py), 8.42 (d, J = 4.7 Hz, 4H, β-pyrrole), 8.39 (d, J = 4.7 Hz, 4H, β-pyrrole), 8.35 (d, J = 4.7 Hz, 4H, β-pyrrole), 8.11 (d, J = 7.3 Hz, 2H, Ph or Py), 7.70 (t, J = 7.9 Hz, 2H, Ph or Py), 7.59 (s, 4H, Ph), 7.19 (s, 4H, mesityl), 7.13 (s, 4H, mesityl), 7.00 (brs, 2H, Ph or Py), 2.59 (s, 12H, CH3), 1.48 (s, 12H, CH3), 1.36 (s, 12H, CH3), −3.44 (s, 4H, inner NH); MALDI-TOF mass: m/z [M + H]+ 1557.6928, calcd. for [C106H85N12O2]+ 1557.6913.

3.5. Synthesis of Fe2-CP2p

In a 10 mL flask were placed Fb2-CP2p (7.1 mg, 4.6 × 10−6 mol), CHCl3 (2.0 mL), and 2,6-lutidine (14 μL, 1.2 × 10−4 mol). Anhydrous FeCl2 (37 mg, 2.9 × 10−4 mol) dissolved in methanol (1.0 mL) was added to it, and the mixture was stirred under reflux for 18 h. The resulting solution was transferred to a perfluoroalkoxy alkanes (PFA)-coated flask, and diluted with CHCl3 (ca. 10 mL). The organic solution was washed with 1 M HCl aqueous solution (×3) and water (×3), and then dried over anhydrous Na2SO4. The residue obtained by evaporation of the solvent was purified with a flush silica gel column (eluents: CHCl3, CHCl3:CH3OH = 50:1, and CHCl3:CH3OH = 10:1). The fraction eluted with CHCl3:CH3OH = 50:1 was collected and the solvent was evaporated. The solid was dissolved in CHCl3 (ca. 10 mL) was treated with 1 M HCl aqueous solution (×3) and water (×3) and passed through Phase Separator paper (Whatman). The solvent was evaporated to dryness, resulting in the titled compound as a black solid (6.3 mg, 79%). MALDI-TOF mass: m/z [M−2Cl + H]+ 1666.5417 (max), calcd. for [C106H81N12O2Fe2]+ 1666.5333; UV-vis absorption (CHCl3) λmax/nm (ε/M−1cm−1) = 375 (7.3 × 104, LMCT band), 416 (1.7 × 105, Soret band), 510 (1.4 × 104, Q band).

3.6. Synthesis of Fe2-P2

In a 10 mL flask were placed 5Fb2 (26 mg, 1.6×10−5 mol), CHCl3 (5.0 mL), and 2,6-lutidine (30 μL, 2.5 × 10−4 mol). Anhydrous FeCl2 (100 mg, 7.9 × 10−4 mol) dissolved in methanol (3.0 mL) was added to it, and the mixture was stirred under reflux for 24 h. The reaction was quenched by adding EDTA aqueous solution, and the resulting solution was transferred to a perfluoroalkoxy alkanes (PFA)-coated flask, and diluted with CHCl3 (ca. 50 mL). The organic solution was washed with water (×3) and brine (×1), and then dried over anhydrous Na2SO4. The use of hydrochloric acid was avoided because the Boc group could be removed under acidic conditions. The residue obtained by evaporation of the solvent was purified with a flush silica gel column (eluents: CHCl3:CH3OH = 50:1 and CHCl3:CH3OH = 10:1). The fraction eluted with CHCl3:CH3OH = 10:1 was collected, and was stirred with brine for overnight. The organic layer was passed through Phase Separator paper (Whatman) and the solvent was evaporated to dryness, giving the titled compound as a black solid (24 mg, 85%). MALDI-TOF mass: m/z [M − 2Boc − 2Cl + H]+ 1534.5110 (max), calcd. for [C98H76Fe2N12]+ 1534.5121; UV-vis absorption (CHCl3) λmax/nm (ε/M−1cm−1) = 378 (8.4 × 104, LMCT band), 421 (1.8 × 105, Soret band), 509 (2.4 × 104, Q band).

3.7. Synthesis of FeP-phen

In a 25 mL flask were placed H2P-phen (30 mg, 3.4 × 10−5 mol), CHCl3 (8.0 mL), and 2,6-lutidine (35 μL, 3.0 × 10−4 mol). Anhydrous FeCl2 (107 mg, 8.5 × 10−4 mol) dissolved in methanol (3.0 mL) was added to it, and the mixture was stirred under reflux for 22 h. The reaction was quenched by adding EDTA (ethylenediaminetetraacetic acid) aqueous solution, and the resulting solution was transferred to a perfluoroalkoxy alkanes (PFA)-coated flask, and diluted with CHCl3 (ca. 50 mL). The organic solution was washed with was washed with 1 M HCl (×3) and brine and then dried over anhydrous Na2SO4. The residue obtained by evaporation of the solvent was purified with a flush silica gel column (eluents: CHCl3, CHCl3:CH3OH = 9:1). The fraction eluted with CHCl3:CH3OH = 9:1 was collected and the solvent was evaporated. The solid was dissolved in CHCl3 (ca. 10 mL) and treated with 1 M HCl aqueous solution (×3) and brine, and passed through Phase Separator paper (Whatman). The solvent was evaporated to dryness, giving the titled compound as a black solid (22 mg, 69%). MALDI-TOF mass: m/z [M − Cl + H]+ 939.3940 (max), calcd. for [C62H55FeN6]+ 939.3833.

3.8. Photocatalytic CO2 Reduction

In glass tubes (8.0 mL, i.d. = 10 mm), 1.0 mL of CO2-saturated DMA solutions containing BIH was added by 1.0 mL of Ar-saturated DMA solutions containing the Fe porphyrin dimer and the photosensitizer, and the reaction solutions were bubbled with CO2 gas (purity ≥ 99.995%) for 15 min. Photo-irradiations were carried out using a merry-go-round irradiation apparatus (Iris-MG, Cell Systems, Yokohama, Japan) equipped with LED lamps at λ = 420 nm (FWHM = 18.4 nm). The gaseous reaction products (CO, H2, and CH4) were quantified with a gas chromatography system (GC-2014, Shimadzu Science, Kyoto, Japan) equipped with a Shincarbon column (i.d. 3.0 mm × 3.0 m) and a thermal conductivity detector (TCD). The product (formate) in the solutions was analyzed with a capillary electrophoresis system (Otuka Electronics Co. CAPI-3300I, Osaka, Japan).

3.9. 13CO2-Labeling Experiment

In glass tubes (8.0 mL, i.d. = 10 mm), 2.0 mL of Ar-saturated DMA solutions containing BIH (0.10 M) was added by 2.0 mL of Ar-saturated DMA solutions containing Zn2-CP2p (10 μM) and Phen2 (1.0 mM), and the reaction solutions were bubbled with 13CO2 gas, which was generated by the addition of 3.0 M sulfuric acid (5 mL) to Ba13CO3 powder (2.5 g). After irradiation at 420 nm (5 mW) for 16 h with the merry-go-round irradiation apparatus, the gaseous reaction products were analyzed with a GC-MS (GCMS-QP2010 Plus, Shimadzu Science; RESTEK (Bellefonte, PA, USA); RT-Msieve 5A). In an NMR tube, a 0.5 mL DMA-d9 solution containing Zn2-CP2p (0.10 mM), Phen2 (1.0 mM) and BIH (0.10 M) was bubbled with 13CO2 gas, which was generated by addition of 3.0 M sulfuric acid (3 mL) to Ba13CO3 powder (1.0 g). After irradiation at 420 nm (5 mW) for 21 h with the merry-go-round irradiation apparatus, the 1H and 13C NMR spectra were measured.

3.10. Computational Methods

The DFT calculations were carried out using the Gaussian 09 package of programs [27]. Each structure was fully optimized using the B3LYP functional using the 6–31G(d) basis set for all atoms, except Fe, and the standard double-ζ type LANL2DZ basis set with the effective core potential of Hay−Wadt for Fe. The stationary points were verified using the vibrational analysis.

4. Conclusions

In this study, we show that the photochemical CO2 reduction yielded CH4, an eight-electron reduction product of CO2, when Fe porphyrins were placed in a face-to-face arrangement. Although the catalytic turnover number of CH4 was low in this study, this can be attributed to the significant degradation of BIH. Thus, it is expected that by addressing this issue, we can enhance the catalytic performance. While bimetallic porphyrin complexes have been reported to enhance CO production, to the best of our knowledge, this is the first report on the induction of CH4 production using bimetallic porphyrins. We anticipate that this finding can contribute to the understanding CH4 formation mechanisms and provide potential molecular design guidelines for selective CH4 generation.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29112453/s1, Figure S1: UV–vis absorption and fluorescence spectra of Zn2-CP2m and Fb2-CP2m; Figure S2: MALDI-TOF mass of Fb2-CP2m; Figure S3: 1H NMR spectrum of Fb2-CP2m; Figure S4: UV–vis absorption and fluorescence spectra of Fb2-CP2m and HCl-treated Fe2-CP2m; Figures S5–S6: MALDI-TOF mass of Fe2-CP2m and Fb2-CP2p; Figure S7: 1H NMR spectrum of Fb2-CP2; Figure S8: UV–vis absorption spectra of Fe2-CP2p before and after treatment with HCl aqueous solution; Figure S9: MALDI-TOF mass of Fe2-CP2p; Figure S10: UV-vis absorption and fluorescence spectra of Fb2-P2 and brine-treated Fe2-P2; Figure S11: MALDI-TOF mass of Fe2-P2; Figure S12: DPVs of Fe2-P2, Fe2-CP2m, and Fe2-CP2p; Figure S13: CVs of Fe2-P2, Fe2-CP2m, and Fe2-CP2p collected in CO2-saturated DMF in the presence of water; Figure S14: CVs of Fe2-P2 and Fe2-CP2m collected in CO2-saturated DMF in the presence of M(OTf)n (M = Mg (n =2), La and Gd (n = 3)); Figure S15: CVs of Fe2-P2 and Fe2-CP2m in CO2-saturated DMF in the presence of M(OTf)n (M = Mg (n =2), La and Gd (n = 3)), and water; Table S1: Effect of each additive on the ratio of the maximum current to ip0; Figure S16: Energy diagram of the photoinduced electron transfer during the formation of Fe(0) species via the oxidative quenching process by either Fe-o-OH or the Fe porphyrin dimer, when Ir(ppy)3 is used as a photosensitizer and TEA as a sacrificial donor; Figure S17: Energy diagram of the photoinduced electron transfer during the formation of Fe(0) species via the reductive quenching process by BIH, when Ir(ppy)3 is used as a photosensitizer and BIH as a sacrificial donor; Figure S18: UV–vis absorption and emission spectra of Ir(ppy)3 in Ar-saturated DMA at 298 K in the presence of various amounts of BIH; Figure S19: Stern–Volmer plot of emission quenching of Ir(ppy)3 by BIH; Table S2: Photocatalytic reaction using Fe-o-OH in acetonitrile; Figure S20: HPLC charts of the reaction solutions containing BIH and Ir(ppy)3 before and after irradiation; Figure S21: Comparison of the reduction products and BIH consumption in Table S2; Figure S22: 1H NMR spectra of the reaction solutions containing BIH and Ir(ppy)3 before and after irradiation; Figure S23: Decomposition of BIH during irradiation in the presence of PS and CO2; Table S3: Effects of solvents and additives on photocatalytic reaction using Fe-o-OH; Figure S24: Comparison of the reduction products and BIH consumption in Table S3; Figure S25: Plots of the amounts of the reduction products and the consumption of BIH after 60 min of irradiation at 450 nm using a merry-go-round apparatus equipped with LED lamps versus the concentration of FeP-phen in CO2-saturated DMA in the presence of Ir(ppy)3 and BIH; Figure S26: Time dependence of the reduction products during the irradiation of CO2-saturated DMA solutions containing FeP-phen in the presence of BIH and Ir(ppy)3 at 450 nm; Figure S27: HPLC charts of the resulting solutions after irradiation in Figure S26, and time dependence of the remaining amount of BIH determined using HPLC; Figure S28: Time dependence of the remaining amount of BIH as per irradiation intensity in blank CO2-saturated DMA solutions containing BIH and Phen2 during irradiation at 420 nm; Figure S29: Dependence of CO production on light intensity at 420 nm in CO2-saturated DMA solutions containing Fe2-P2, BIH, and Phen2; Figure S30: TONs of CO, H2, and CH4 during irradiation at 420 nm for 18 h in CO2-saturated DMA solutions containing Fe2-CP2p, BIH, and Phen2 in the presence of acids; Figure S31: UV–vis absorption and fluorescence spectra of Phen2 in Ar-saturated DMA in the presence of various amounts of BIH; Figure S32: Stern–Volmer plot of emission quenching of Phen2 by BIH, Figure S33: Relationship between the reduction products and the initial concentration of BIH during irradiation at 420 nm for 4 h and 18 h in CO2-saturated DMA solutions containing Fe2-CP2p and Phen2; Figure S34: Time dependence of the reduction products during the irradiation of CO2-saturated DMA solutions containing Fe2-CP2p in the presence of BIH and Phen2 at 420 nm, and gas chromatogram of the gaseous reaction products after the irradiation for 35 h; Figure S35: Gas chromatograms of the resulting gas-phase products after irradiation at 420 nm for 14 h under 12CO2 or 13CO2 atmosphere in DMA solutions containing Fe2-CP2p, BIH, and Phen2 obtained using mass spectroscopy; Figure S36: Mass spectra of CH4 generated under 12CO2 and 13CO2 atmosphere, and gas chromatogram of the products obtained from the reaction under 13CO2 atmosphere, plotted for each m/z; Figure S37: 1H NMR spectra of the reaction solutions containing Fe2-CP2p, BIH, and Phen2 during irradiation at 420 nm under a 13CO2 atmosphere; Figure S38: Comparison of the 1H NMR spectra of reaction solutions containing Fe2-CP2p, BIH, and Phen2 after irradiation at 420 nm for 21 h under 12CO2 and 13CO2 atmospheres; Figure S39: 13C NMR spectra of the reaction solutions containing Fe2-CP2p, BIH, and Phen2 during irradiation at 420 nm under a 13CO2 atmosphere; Figure S40: HMBC of the reaction solutions containing Fe2-CP2p, BIH, and Phen2 during irradiation for 21 h at 420 nm under 12CO2 and 13CO2 atmospheres; Figure S41: HSQC of the reaction solutions containing Fe2-CP2p, BIH, and Phen2 during irradiation for 21 h at 420 nm under a 13CO2 atmosphere; Figure S42: 1H-1H COSY of the reaction solutions containing Fe2-CP2p, BIH, and Phen2 during irradiation for 21 h at 420 nm under a 13CO2 atmosphere. Scheme S1. Synthetic routes of Fe2-CP2m, p. Scheme S2. Synthetic route of Fe2-P2. Scheme S3. Synthetic route of FeP-phen.

Author Contributions

Conceptualization, Y.K.; formal analysis, Y.K., M.H., and A.S.; investigation, M.H.; writing—original draft preparation, Y.K.; writing—review and editing, A.S.; supervision, A.S.; project administration, Y.K.; funding acquisition, Y.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by ENEOS Hydrogen Trust Fund and JSPS KAKENHI (grant number: JP22H02186).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article/Supplementary Materials, and further inquiries can be directed to the corresponding author/s.

Acknowledgments

We thank Shunya Yoshino, Yuichi Yamaguchi, and Akihiko Kudo (Tokyo University of Science) for help with the GC-MS and capillary electrophoresis analyses and Hiroyuki Koshino (RIKEN) for help with the NMR analyses.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Sato, S.; Arai, T.; Morikawa, T.; Uemura, K.; Suzuki, T.M.; Tanaka, H.; Kajino, T. Selective CO2 Conversion to Formate Conjugated with H2O Oxidation Utilizing Semiconductor/Complex Hybrid Photocatalysts. J. Am. Chem. Soc. 2011, 133, 15240. [Google Scholar] [CrossRef]
  2. Sahara, G.; Kumagai, H.; Maeda, K.; Kaeffer, N.; Artero, V.; Higashi, M.; Abe, R.; Ishitani, O. Photoelectrochemical Reduction of CO2 Coupled to Water Oxidation Using a Photocathode with a Ru(II)–Re(I) Complex Photocatalyst and a CoOx/TaON Photoanode. J. Am. Chem. Soc. 2016, 138, 14152. [Google Scholar] [CrossRef] [PubMed]
  3. Costentin, C.; Robert, M.; Savéant, J.-M. Current Issues in Molecular Catalysis Illustrated by Iron Porphyrins as Catalysts of the CO2-to-CO Electrochemical Conversion. Acc. Chem. Res. 2015, 48, 2996. [Google Scholar] [CrossRef] [PubMed]
  4. Takeda, H.; Cometto, C.; Ishitani, O.; Robert, M. Electrons, Photons, Protons and Earth-Abundant Metal Complexes for Molecular Catalysis of CO2 Reduction. ACS Catal. 2017, 7, 70. [Google Scholar] [CrossRef]
  5. Costentin, C.; Drouet, S.; Robert, M.; Savéant, J.-M. A Local Proton Source Enhances CO2 Electroreduction to CO by a Molecular Fe Catalyst. Science 2012, 338, 90. [Google Scholar] [CrossRef]
  6. Mohamed, E.A.; Zahran, Z.N.; Naruta, Y. Efficient electrocatalytic CO2 reduction with a molecular cofacial iron porphyrin dimer. Chem. Commun. 2015, 51, 16900. [Google Scholar] [CrossRef]
  7. Zhang, C.; Gotico, P.; Guillot, R.; Dragoe, D.; Leibl, W.; Halime, Z.; Aukauloo, A. Bio-Inspired Bimetallic Cooperativity through a Hydrogen Bonding Spacer in CO2 Reduction. Angew. Chem. Int. Ed. 2023, 62, e202214665. [Google Scholar] [CrossRef]
  8. Azcarate, I.; Costentin, C.; Robert, M.; Savéant, J.-M. Through-Space Charge Interaction Substituent Effects in Molecular Catalysis Leading to the Design of the Most Efficient Catalyst of CO2-to-CO Electrochemical Conversion. J. Am. Chem. Soc. 2016, 138, 16639. [Google Scholar] [CrossRef] [PubMed]
  9. Rao, H.; Schmidt, L.C.; Bonin, J.; Robert, M. Visible-light-driven methane formation from CO2 with a molecular iron catalyst. Nature 2017, 548, 74. [Google Scholar] [CrossRef]
  10. Rao, H.; Lim, C.-H.; Bonin, J.; Miyake, G.M.; Robert, M. Visible-Light-Driven Conversion of CO2 to CH4 with an Organic Sensitizer and an Iron Porphyrin Catalyst. J. Am. Chem. Soc. 2018, 140, 17830. [Google Scholar] [CrossRef]
  11. Chen, J.Y.; Li, M.; Liao, R.Z. Mechanistic Insights into Photochemical CO2 Reduction to CH4 by a Molecular Iron–Porphyrin Catalyst. Inorg. Chem. 2023, 62, 9400. [Google Scholar] [CrossRef] [PubMed]
  12. Hashimoto, M.; Kuramochi, Y.; Ito, S.; Kinbara, Y.; Satake, A. Metal-templated synthesis of rigid and conformationally restricted cyclic bisporphyrins: Specific retention times on a cyanopropyl-modified silica gel column. Org. Biomol. Chem. 2021, 19, 3159. [Google Scholar] [CrossRef] [PubMed]
  13. Tamaki, Y.; Koike, K.; Morimoto, T.; Ishitani, O. Substantial improvement in the efficiency and durability of a photocatalyst for carbon dioxide reduction using a benzoimidazole derivative as an electron donor. J. Catal. 2013, 304, 22. [Google Scholar] [CrossRef]
  14. Aranzaes, J.R.; Daniel, M.-C.; Astruc, D. Metallocenes as references for the determination of redox potentials by cyclic voltammetry—Permethylated iron and cobalt sandwich complexes, inhibition by polyamine dendrimers, and the role of hydroxy-containing ferrocenes. Can. J. Chem. 2006, 84, 288. [Google Scholar] [CrossRef]
  15. Hammouche, M.; Lexa, D.; Momenteau, M.; Savéant, J.-M. Chemical catalysis of electrochemical reactions. Homogeneous catalysis of the electrochemical reduction of carbon dioxide by iron(“0”) porphyrins. Role of the addition of magnesium cations. J. Am. Chem. Soc. 1991, 113, 8455. [Google Scholar] [CrossRef]
  16. Bhugun, I.; Lexa, D.; Savéant, J.-M. Catalysis of the electrochemical reduction of carbon dioxide by iron (0) porphyrins. Synergistic effect of Lewis acid cations. J. Phys. Chem. 1996, 100, 19981. [Google Scholar] [CrossRef]
  17. Anxolabéhère-Mallart, E.; Bonin, J.; Fave, C.; Robert, M. Small-molecule activation with iron porphyrins using electrons, photons and protons: Some recent advances and future strategies. Dalton Trans. 2019, 48, 5869. [Google Scholar] [CrossRef] [PubMed]
  18. Kuramochi, Y.; Ishitani, O.; Ishida, H. Reaction mechanisms of catalytic photochemical CO2 reduction using Re (I) and Ru (II) complexes. Coord. Chem. Rev. 2018, 373, 333. [Google Scholar] [CrossRef]
  19. Kuramochi, Y.; Kamiya, M.; Ishida, H. Photocatalytic CO2 Reduction in N,N-Dimethylacetamide/Water as an Alternative Solvent System. Inorg. Chem. 2014, 53, 332. [Google Scholar] [CrossRef]
  20. Pearson, R.M.; Lim, C.-H.; McCarthy, B.G.; Musgrave, C.B.; Miyake, G.M. Organocatalyzed Atom Transfer Radical Polymerization Using N-Aryl Phenoxazines as Photoredox Catalysts. J. Am. Chem. Soc. 2016, 138, 11399. [Google Scholar] [CrossRef]
  21. Kuramochi, Y.; Suzuki, Y.; Asai, S.; Suzuki, T.; Iwama, H.; Asano, M.S.; Satake, A. Significance of the connecting position between Zn(II) porphyrin and Re(I) bipyridine tricarbonyl complex units in dyads for room-temperature phosphorescence and photocatalytic CO2 reduction: Unexpected enhancement by triethanolamine in catalytic activity. Chem. Sci. 2023, 14, 8743. [Google Scholar] [CrossRef] [PubMed]
  22. McCarthy, B.G.; Pearson, R.M.; Lim, C.-H.; Sartor, S.M.; Damrauer, N.H.; Miyake, G.M. Structure–property relationships for tailoring phenoxazines as reducing photoredox catalysts. J. Am. Chem. Soc. 2018, 140, 5088. [Google Scholar] [CrossRef] [PubMed]
  23. Lim, C.-H.; Ilic, S.; Alherz, A.; Worrell, B.T.; Bacon, S.S.; Hynes, J.T.; Glusac, K.D.; Musgrave, C.B. Benzimidazoles as Metal-Free and Recyclable Hydrides for CO2 Reduction to Formate. J. Am. Chem. Soc. 2018, 141, 272. [Google Scholar] [CrossRef] [PubMed]
  24. Xie, W.; Xu, J.; Idros, U.M.; Katsuhira, J.; Fuki, M.; Hayashi, M.; Yamanaka, M.; Kobori, Y.; Matsubara, R. Metal-free reduction of CO2 to formate using a photochemical organohydride-catalyst recycling strategy. Nat. Chem. 2023, 15, 794. [Google Scholar] [CrossRef] [PubMed]
  25. Salzmann, R.; Ziegler, C.J.; Godbout, N.; McMahon, M.T.; Suslick, K.S.; Oldfield, E. Carbonyl Complexes of Iron(II), Ruthenium(II), and Osmium(II) 5,10,15,20-Tetraphenylporphyrinates:  A Comparative Investigation by X-ray Crystallography, Solid-State NMR Spectroscopy, and Density Functional Theory. J. Am. Chem. Soc. 1998, 120, 11323. [Google Scholar] [CrossRef]
  26. Kuramochi, Y.; Fujisawa, Y.; Satake, A. Photocatalytic CO2 Reduction Mediated by Electron Transfer via the Excited Triplet State of Zn(II) Porphyrin. J. Am. Chem. Soc. 2020, 142, 705. [Google Scholar] [CrossRef]
  27. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 09, revision D.01; Gaussian, Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
Figure 1. Chemical structures of Fe porphyrins (molecular models optimized by density functional theory calculations using B3LYP/LANL2DZ/6–31G(d) for the Fe porphyrin dimers).
Figure 1. Chemical structures of Fe porphyrins (molecular models optimized by density functional theory calculations using B3LYP/LANL2DZ/6–31G(d) for the Fe porphyrin dimers).
Molecules 29 02453 g001
Figure 2. (Left) Interaction of transition dipole moments of the bisporphyrin and UV−vis absorption spectra of FeTPP(Cl) (black dotted line; TPP = tetraphenylporphyrin), Fe2-P2 (blue line), Fe2-CP2m (red line), Fe2-CP2p (green line) in CHCl3. The inset shows the magnification of the Soret band region: (right) CVs (scan rate = 100 mV s−1) of (a) Fe2-P2, (b) Fe2-CP2m, and (c) Fe2-CP2p (0.3 mM) collected in dry DMF under Ar (solid lines) and CO2 (dotted lines) atmospheres with 0.1 M nBu4NPF6 as the supporting electrolyte.
Figure 2. (Left) Interaction of transition dipole moments of the bisporphyrin and UV−vis absorption spectra of FeTPP(Cl) (black dotted line; TPP = tetraphenylporphyrin), Fe2-P2 (blue line), Fe2-CP2m (red line), Fe2-CP2p (green line) in CHCl3. The inset shows the magnification of the Soret band region: (right) CVs (scan rate = 100 mV s−1) of (a) Fe2-P2, (b) Fe2-CP2m, and (c) Fe2-CP2p (0.3 mM) collected in dry DMF under Ar (solid lines) and CO2 (dotted lines) atmospheres with 0.1 M nBu4NPF6 as the supporting electrolyte.
Molecules 29 02453 g002
Figure 3. Time dependence of the reduction products generated during the irradiation of CO2-saturated DMA solutions (2.0 mL) containing (a) Fe2-P2 and (b) Fe2-CP2m (10 μM) in the presence of BIH (10 mM) and Ir(ppy)3 (0.2 mM) at 450 nm using a merry-go-round apparatus equipped with LED lamps (input power: 10 mW).
Figure 3. Time dependence of the reduction products generated during the irradiation of CO2-saturated DMA solutions (2.0 mL) containing (a) Fe2-P2 and (b) Fe2-CP2m (10 μM) in the presence of BIH (10 mM) and Ir(ppy)3 (0.2 mM) at 450 nm using a merry-go-round apparatus equipped with LED lamps (input power: 10 mW).
Molecules 29 02453 g003
Figure 4. Time dependence of the reduction products generated during the irradiation of CO2-saturated DMA solutions (2.0 mL) containing (a) Fe2-P2, (b) Fe2-CP2m, and (c) Fe2-CP2p (10 μM) in the presence of BIH (10 mM) and Phen2 (1 mM) at 420 nm using a merry-go-round apparatus equipped with LED lamps (Input power: 5 mW).
Figure 4. Time dependence of the reduction products generated during the irradiation of CO2-saturated DMA solutions (2.0 mL) containing (a) Fe2-P2, (b) Fe2-CP2m, and (c) Fe2-CP2p (10 μM) in the presence of BIH (10 mM) and Phen2 (1 mM) at 420 nm using a merry-go-round apparatus equipped with LED lamps (Input power: 5 mW).
Molecules 29 02453 g004
Figure 5. Relationship between the reduction products and the initial concentration of BIH during irradiation at 420 nm for 4 h in CO2-saturated DMA solutions (2.0 mL) containing Fe2-CP2p (10 μM) and Phen2 (1 mM), using a merry-go-round apparatus equipped with LED lamps (input power: 5 mW).
Figure 5. Relationship between the reduction products and the initial concentration of BIH during irradiation at 420 nm for 4 h in CO2-saturated DMA solutions (2.0 mL) containing Fe2-CP2p (10 μM) and Phen2 (1 mM), using a merry-go-round apparatus equipped with LED lamps (input power: 5 mW).
Molecules 29 02453 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kuramochi, Y.; Hashimoto, M.; Satake, A. Methane Formation Induced via Face-to-Face Orientation of Cyclic Fe Porphyrin Dimer in Photocatalytic CO2 Reduction. Molecules 2024, 29, 2453. https://doi.org/10.3390/molecules29112453

AMA Style

Kuramochi Y, Hashimoto M, Satake A. Methane Formation Induced via Face-to-Face Orientation of Cyclic Fe Porphyrin Dimer in Photocatalytic CO2 Reduction. Molecules. 2024; 29(11):2453. https://doi.org/10.3390/molecules29112453

Chicago/Turabian Style

Kuramochi, Yusuke, Masaya Hashimoto, and Akiharu Satake. 2024. "Methane Formation Induced via Face-to-Face Orientation of Cyclic Fe Porphyrin Dimer in Photocatalytic CO2 Reduction" Molecules 29, no. 11: 2453. https://doi.org/10.3390/molecules29112453

Article Metrics

Back to TopTop