Next Article in Journal
In Silico Exploration of Natural Antioxidants for Sepsis Drug Discovery
Previous Article in Journal
Chitosan-Oligosaccharide-Bearing Biphasic Calcium Phosphate Bone Cement: Preparation and Angiogenic Activity In Vitro
Previous Article in Special Issue
Tailoring Catalysts for CO2 Hydrogenation: Synthesis and Characterization of NH2–MIL–125 Frameworks
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Progress in Heteroatom-Containing Metalloporphyrin-Based Catalysts for CO2 Reduction

1
Key Laboratory of Cluster Science of Ministry of Education, Beijing Key Laboratory of Photoelectronic/Electrophotonic Conversion Materials, School of Chemistry and Chemical Engineering, Beijing Institute of Technology, Beijing 100081, China
2
School of Chemistry, Chemical Engineering and Materials, Jining University, Qufu 273155, China
*
Authors to whom correspondence should be addressed.
Molecules 2025, 30(11), 2287; https://doi.org/10.3390/molecules30112287
Submission received: 24 April 2025 / Revised: 18 May 2025 / Accepted: 20 May 2025 / Published: 23 May 2025
(This article belongs to the Special Issue Innovative Chemical Pathways for CO2 Conversion)

Abstract

:
Metalloporphyrins, owing to their structural resemblance to natural enzyme active sites and highly tunable coordination environments, have emerged as promising catalysts for converting CO2 into value-added chemicals and fuels. Considerable efforts have been made to modify metalloporphyrins to improve their catalytic capability for CO2 reduction. One approach involves modifying the metal coordination environment (known as the first coordination sphere) to generate heteroatom-containing metalloporphyrins, particularly N-confused and O/S-substituted variants. While heteroatom-containing metalloporphyrins were first synthesized in 1989, their use in CO2 reduction catalysis was not reported until after 2020. Herein, we review the recent progress in the design, catalytic performance, and mechanistic studies of N-confused and O/S-substituted metalloporphyrins towards CO2 reduction. This review encompasses both experimental and theoretical computational work, as well as the use of porphyrins as catalysts in photocatalysis and electrocatalysis. Finally, based on the current research advances, we present critical recommendations and future research directions, with a focus on theoretical studies, in the hope of facilitating the rational design of novel catalysts for sustainable energy conversion and environmental remediation.

Graphical Abstract

1. Introduction

The increasing concerns about climate change and the depletion of fossil fuels have raised the need for sustainable energy solutions [1,2]. The electrochemical reduction of carbon dioxide (CO2) to value-added hydrocarbons and chemical feedstocks, such as alcohols, aldehydes, ketones, etc., represents a promising approach for dealing with CO2 emissions and alleviating energy and environmental crisis [3,4,5]. However, the inherent chemical inertness of CO2 and the existence of a competitive hydrogen evolution reaction poses a challenge for its efficient conversion [6,7,8], thereby driving the need for catalyst development [9,10].
Metalloporphyrins are a class of coordination complexes in which a metal ion (e.g., Fe, Mg, Co, Zn, Cu, Ni) is bound to the nitrogen atoms of a porphyrin ligand. They are ubiquitous in nature and serve essential functions in various critical life processes. Typical examples include iron porphyrins (e.g., heme and cytochrome P450), magnesium porphyrins (e.g., chlorophyll), and cobalt porphyrins (e.g., vitamin B12) [11,12]. These metalloporphyrins share a common mechanistic hallmark: enabling electron transfer via redox transitions of the central metal ion (e.g., Fe2+/Fe3+) or activating small molecules (e.g., O2, CO2) through metal–ligand coordination [13,14]. Their unparalleled catalytic versatility has driven breakthroughs in artificial enzyme design, with transformative applications in precision medicine like targeted drug metabolism and renewable energy, particularly CO2-to-fuel conversion [15,16,17]. Regarding the latter, much empirical evidence has been reported in the literature. For example, iron porphyrin compounds can act as homogeneous catalysts for the reduction of CO2 to CO in N, N-dimethylformamide (DMF) solution [18]. Nickel porphyrin compounds featuring exocyclic 1,3-ditert-butylimidazolin-2-ylidenamino (NItBu) substituents also exhibit catalytic activity for CO2-to-CO conversion in DMF [19]. Additionally, cobalt porphyrin compounds can be incorporated into metal–organic frameworks (MOFs) or covalent–organic frameworks (COFs) through organic linkers, serving as efficient catalysts for CO2 reduction to CO [20,21,22,23,24,25].
How to optimize the catalytic activity and selectivity of the metalloporphyrin-based catalysts is the key to these studies. To reach this end, modifying the structures of the metalloporphyrins has been shown to be an effective strategy [26,27]. The distinctive structure of porphyrin molecules provides a highly modifiable platform for functionalization. As shown in Figure 1a, porphyrins are macrocycles that are composed of four pyrrole subunits interconnected via methine bridges, forming a planar, aromatic, and highly stable framework [28]. It readily anchors metal ions at its core, forming a stable M–N4 coordination environment (Figure 1b) [29]. In addition to changing the type of the central metal, common strategies for modifying metal porphyrins include introducing substituents at the meso or β positions of the porphyrin, coordinating axial ligands to the central metal, and modifying the first coordination or the outer coordination sphere [27,30,31,32,33,34]. Among these strategies, modifying the first coordination sphere of metal porphyrin compounds has received relatively less attention, likely due to the greater synthetic challenges in constructing asymmetrically coordinated metal porphyrin systems. Nevertheless, recent experimental and theoretical studies have made significant progress in this direction, demonstrating it as an effective strategy for enhancing catalytic efficiency [27,35,36,37]. Of note, this strategy has been widely used in graphene-based M–N–C type single-atom catalysts [38,39,40,41,42].
A number of recent review articles have provided comprehensive coverage of utilizing metal porphyrin-based catalysts for CO2 conversion [14,26,43,44,45]. These reviews all focus on traditional N4-metalloporphyrin systems. For example, Philipp Gotico et al. [44] (2020) and Domingo-Tafalla et al. [26] (2023) reviewed research advances in outer coordination sphere modifications through functional group effects (e.g., hydrogen bonding, electrostatic interactions, and electronic effects) and hierarchical material design (e.g., MOFs, COFs, and carbon composites) of metalloporphyrins. To our knowledge, no review has addressed modifications of the first coordination sphere in metal porphyrins for CO2 reduction, motivating our summary of recent advances in this field.
Modifying the first coordination sphere in metal porphyrins enables access to diverse derivatives with potentially improved catalytic performance. For example, N-confused porphyrins, which are isomers of porphyrins, are formed by inverting the N and C atoms of one or multiple pyrrole rings (Figure 1c). Additionally, replacing one or two pyrrole N atoms with other heteroatoms such as O, S, Se, B, P, and C results in a class of porphyrinoids known as heteroatom-containing porphyrins or core-modified porphyrins (Figure 1d) [46,47,48]. Among the latter, metalloporphyrins incorporating oxygen or sulfur heteroatoms demonstrate relatively facile synthesis. These substitutions induce significant modifications in both the geometric and electronic structures of the metalloporphyrins, and they fundamentally change the reaction coordinate and introduce additional mechanistic complexity in CO2 reduction catalytic pathways.
Herein, we present a review of the advancements of first coordination sphere modification on metalloporphyrins for CO2 reduction, both experimentally and theoretically. We particularly focus on the N-confused and O- or S-substituted species. The contents of this review are organized as follows: Session 2 briefly reviews the application of metalloporphyrins in CO2 reduction and introduces the commonly used computational methods in modeling this process. Session 3 reviews the synthesis and catalytic progress of N-confused metalloporphyrin for CO2RR. Session 4 reviews the synthesis and catalytic progress of O/S-substituted metalloporphyrin for CO2RR. Session 5 presents a summary and outlook. By reviewing the current state of research on first coordination modification on metalloporphyrins, we hope to provide insights into the design principles and strategies that can guide the development of next-generation catalysts for sustainable energy conversion.

2. Metalloporphyrins for CO2RR

The use of metalloporphyrins as catalysts for CO2 reduction can be traced back to 1979. Toshima et al. found that cobalt porphyrins with carboxyl (–COOH) and sulfonic acid groups (–SO3H) showed electrocatalytic activity in reducing CO2 to formic acid, HCOOH [49]. Since then, metalloporphyrins have garnered increasing attention as catalysts for CO2 reduction, especially those with low-cost transition metal centers like iron, cobalt, and nickel [18,19,43,50,51]. For example, Savéant et al. (2016) reported electrochemical CO2-to-CO conversion by Fe tetraphenylporphyrins with four positively charged trimethylanilinium groups (Fe-p-TMA) [33]. The Omar Yaghi group (2015) synthesized co-porphyrin covalent organic frameworks (COF-366-Co) for the electrocatalytic reduction of CO2 to CO [20]. The Koper group (2015) immobilized cobalt porphyrins on pyrolyzed graphene as electrocatalysts to reduce CO2 to CO and CH4 [52]. Dey’s group (2021) achieved the conversion of CO2 to HCOOH via electroreduction by introducing a pendant amine group onto the iron porphyrin (Fe-2,3,7,8-tetracetyl-15-phenylporphyrin, FeTEsP) [53]. Liu et al. (2013) reported copper porphyrin-based metal–organic frameworks (Cu-5,10,15,20-tetrakis(4-carboxyphenyl) porphyrin, Cu-TCPP-MOFs) as photocatalysts for converting CO2 to CH3OH [54].
The above studies indicate that metal porphyrin compounds predominantly catalyze the reduction of CO2 to C1 products, including carbon monoxide (CO) [18,20,31,33,50,55], formic acid (HCOOH) [49,53], methanol (CH3OH) [54], and methane (CH4) [52,56,57]. C2 products are scarcely observed on porphyrin-based catalysts because of the lack of suitable adsorption sites for C–C coupling to C2 intermediates. Hence, modifications to the coordination sphere of metal porphyrins retain this selectivity, with C1 species—particularly CO—remaining the dominant reduction products.
Elucidating the precise mechanism of CO2 reduction by metalloporphyrin compounds faces several fundamental challenges: the complexity of possible reaction pathways sensitive to experimental conditions, the transient nature of key intermediates limiting experimental detection, and ambiguous electronic configurations arising from redox-active metal centers and porphyrin ligands [58,59]. Hence, a computational study plays an important role. In theoretical calculations, density functional theory (DFT) is almost exclusively employed. Researchers have adopted two distinct modeling approaches, molecular cluster-based DFT calculations and periodic DFT calculations. Molecular cluster models are used for molecular metalloporphyrin complexes as homogeneous catalysts. Electron transfer (ET) and proton transfer (PT) processes are typically treated separately [56,60,61,62,63,64,65,66], enabling the computation of reduction potentials for different species and identifying the order and/or synchronicity of ET and PT. The periodic DFT calculations, which are widely used for heterogeneous catalysts, are primarily applied to covalent organic frameworks (COFs) and metal–organic frameworks (MOFs) constructed with metalloporphyrin units [37,67,68]. Periodic calculations generally consider concerted proton-electron transfer (CPET) as a unified process, as shown in Figure 2. In this case, for example, the reduction of CO2 to CO follows the mechanism of the adsorption of CO2, the formation of *COOH and *CO, and the desorption of CO. The computational hydrogen electrode (CHE) model [69] was usually used to account for the energy change in adding a proton and an electron, i.e., a hydrogen atom. The effects of pH and applied potential can be incorporated through an energy correction term [70].

3. N-Confused Metalloporphyrin

N-confused porphyrins (NCPs) are isomers of porphyrin. The first NCP, 2-aza-21-carbaporphyrin, was synthesized by Latos-Grażyński and Furuta et al. in 1994 [72]. In principle, inverting one, two, or three pyrrole rings of porphyrins could generate a first coordination sphere with N3C1, N2C2, or N1C3 configurations. To date, synthetic efforts reported in the experimental literature have exclusively produced configurations featuring either single or double pyrrole ring inversion, while no successful construction of triply inverted configurations has been achieved. For this reason, theoretical studies mainly focused on N3C1 and N2C2 configurations.
The formal oxidation state of the central metal in a M–N4-type metalloporphyrin is +2. Anchoring metal elements into NCPs gives N-confused metalloporphyrins, and the central metal remains a formal oxidation state of +2. The first example of N3C1-type metalloporphyrins was A Ni(II) complex of 21-(CH3)–NCTPP and was reported by Latos-Grażyński et al. in 1995 [73]. Later, in 2000, the same group reported the synthesis of CuII–N(–CH3)CTPP [74]. FeII(–Br)–NCTPP (2001) and MnII(–Br)–NCTPP (2002) were synthesized by Hung et al. [75,76]. CuII/AgIII/PdII–NCP–(C6F5)4 was synthesized by Furuta et al. in 2003 [77], and CoIINCTPP was synthesized by Bettelheim et al. in 2004 [78]. Other than these, N3C1-type metalloporphyrins with central metals of Mo (2005) [79], Rh (2008) [80], Re (2007) [81], Pt (2000) [82], Au (2005) [83], and Hg (2008) [84] have also been synthesized experimentally. Shao and coworkers (2025) recently reported the synthesis of NiN3C- and CoN3C-coordinated porphyrin-based COFs [85]. For N2C2-type metalloporphyrins, the syntheses of complexes with two inverted carbon atoms both in the ortho-(cis-N2CP) and para-positions (trans-N2CP) have been reported. CuIII/AgIIIcis-N2CP–(C6F5)4 and CuIIItrans-N2CP–(C6F5)4 were synthesized by Furuta et al. in 2000 and 2003 [86,87], respectively. These N-confused metalloporphyrins are illustrated in Figure 3. Computational studies on these systems with other central metals, such as Co, have been conducted [71].
The application of N-confused metalloporphyrins as catalysts for CO2 reduction has only been demonstrated recently. In 2023, Ren and coworkers reported theoretical studies of N-confused metalloporphyrin-based covalent organic frameworks (Por-COFs) for enhanced electrocatalytic CO2RR, where periodic DFT with PBE-D3 functional was used [71]. The study employed a monolayer COF model constructed through imine-linked 5,10,15,20-tetrakis(4-benzaldehyde)porphyrin (Por-CHO) and p-phenylenediamine (PPDA), forming a 2D periodic structure (Figure 4b). Through comprehensive screening of ten 3d transition metals (Sc-Zn) based on stability and catalytic selectivity and activity, CoN4-Por-COF was identified as a promising candidate for catalyzing CO2 reduction to CO (limiting potential UL = −0.89 V), while CrN4-Por-COF catalyzes CO2 reduction to HCOOH (UL = −0.78 V). The authors further modified the coordination environment of CoN4-Por-COF by inverting one or two pyrrole units to design mono-pyrrole or dipyrrole-inverted N-confused Por-COFs, named CoN3C1- and CoN2C2-Por-COFs (Figure 4a). A free energy analysis (ΔG) of key intermediates (*COOH vs. *H) revealed the designed N-confused materials effectively suppressed hydrogen evolution (HER) competition. Detailed calculation of the CO2 reduction pathway (CO2 → COOH → CO → further reduction) showed that the N-confused CoN3C1- and CoN2C2-Por-COFs display improved performance in catalyzing CO2RR than parent CoN4-Por-COF. The limiting potential of reducing CO2 to CO decreases from −0.89 V (CoN4) to −0.76 V (CoN3C1) and −0.60 V (CoN2C2), and applying high potential can yield deep-reduction degree C1 products such as CH3OH and CH4 (Figure 4c–e). Electronic structure analysis revealed that substituting CoN4 to CoN3C1/CoN2C2 increases the electron density on the Co atom and raises the d-band center, thus stabilizing the key intermediates *COOH and lowering the ΔG of respective PDSs (PDS refers to potential-determining step). These findings underscore the promise of coordination engineering in COFs, positioning CoN2C2 and CrN3C1 as superior candidates for selective CO2RR. This theoretical framework not only rationalizes experimental trends but also sets the stage for exploring asymmetric coordination motifs in other transition-metal systems.
Later, Huang and coworkers (2024) [88] synthesized a cobalt porphyrin complex with N-confused coordination, and Shao and coworkers (2025) [85] synthesized a CoN3C porphyrin-based COF. Both works demonstrated that the CoN3C systems enhance electrocatalytic activity for the reduction of O2 to H2O. Although the experiments were not for CO2RR, the authors confirmed the N4 to N3C coordination change reduced electron transfer from cobalt to the porphyrin ligand, which is the key for catalytic performance improvements, consistent with the theoretical prediction by Ren’s work [71]. While the periodic DFT approach effectively modeled the electronic structure of monolayer COFs, the study [71] acknowledged limitations including unaccounted interlayer interactions, solvent effects, and kinetic barriers that might lead to overestimated catalytic performance in practical applications. By assuming a CPET process, the kinetic barriers were not estimated in this periodic DFT work.
In 2024, Hua and coworkers reported a N-confused copper(II) tetraphenylporphyrin (CuNCP) complex for efficient electrochemical CO2 reduction to methane (CH4) under pulsed potential electrolysis (PPE) [89]. Two copper porphyrin complexes were synthesized via a modified Geier method: symmetric copper tetraphenylporphyrin (CuTPP, Cu-N4 coordination) and asymmetric N-confused copper porphyrin (CuNCP, Cu-N3C coordination). Experimental results revealed that CuTPP maintained structural integrity under constant potential electrolysis (CPE) but primarily produced hydrogen (H2) (Figure 4b). CuNCP exhibited higher CO2 reduction activity (Figure 5a), which is attributed to the asymmetric N3C coordination but suffers from structural instability during electrolysis. In situ X-ray absorption spectroscopy (XAS) and transmission electron microscopy (TEM) confirmed the irreversible reduction in CuNCP to Cu0 nanoparticles under CPE, leading to diverse products (CO, C2H4, H2, etc.). However, applying PPE (anodic potential: +1.3 V, cathodic potential: −1.6 V, pulse cycle: 5 s/10 s) significantly enhanced CH4 selectivity, achieving 60% Faradaic efficiency (FE) with a partial current density of 170 mA cm−2 (Figure 5c,d). Combined in situ infrared (IR), Raman spectroscopy, and scanning electrochemical microscopy (SECM) analyses suggested that PPE promotes water dissociation at high anodic bias, consuming OH and lowering local pH, thereby enriching proton availability for the successive hydrogenation of *CO intermediates and ultimately boosting methane production. Although initial hypotheses proposed dynamic Cu(I)/Cu(II) cycling to restore molecular structure, experimental data ruled out catalyst regeneration, instead highlighting the role of electrolysis protocols in optimizing interfacial chemical environments. This study underscores the trade-off between enhanced catalytic activity and compromised stability in asymmetric coordination (CuN3C), and it proposes a novel strategy of tailoring electrolytic conditions to catalyst properties for efficient CO2 reduction.
In 2024, Peng and coworkers reported a nickel(II) N-confused tetraphenylporphyrin complex (NiNCP) (Figure 6a) that efficiently photocatalyzes CO2 to CO reduction in acetonitrile solution under irradiation from a blue LED (420 nm), using [Ru(bpy)3]2+ as a photosensitizer and triisopropanolamine (TIPA) as a sacrificial electron donor [36]. The complex achieved a turnover number (TON) of up to 217,000 with 98% selectivity (Figure 6b). In their experiments, the authors synthesized NiNCP by improving upon the Lindsey and Choi method, confirming its structure through characterization techniques such as MALDI-TOF, XPS, and 1H NMR. They discovered that the acidic N–H group in NiNCP is deprotonated by TIPA during the catalytic process, forming an anionic active species, NiNCP. Photocatalytic experiments revealed that NiNCP produced 0.32 mmol of CO within the first hour of the reaction, and the TON increased significantly as the catalyst concentration decreased (Figure 5b). Through steady-state fluorescence and time-resolved spectroscopy analyses, the authors found that NiNCP efficiently extracts electrons from the photosensitizer, with its LUMO energy level (3.40 eV) well matched to those of the photosensitizer (3.19 eV) and the CO2/CO couple (3.91 eV) (Figure 6c). Furthermore, DFT calculations indicated that NiNCP attacks the carbon atom of CO2 via a ligand-centered Lewis basic site, forming a stable *COOH intermediate stabilized by a strong C–C bond (1.5 Å), rather than following the traditional metal-centered activation pathway, thereby significantly reducing the reaction energy barrier (Figure 6d,e). The work redefines the role of coordination asymmetry, demonstrating that ligand participation—not just metal tuning—can unlock unprecedented catalytic efficiency.
To our knowledge, only the aforementioned studies have been reported on N-confused porphyrins for CO2 research. However, additional works exist regarding metallated NCP complexes applied to the oxygen reduction reaction [85,88]. In the next section, we will discuss the O/S-substituted metalloporphyrins.

4. O/S-Substituted Metalloporphyrin

Since one oxygen or sulfur atom is equivalent to one NH group, directly substituting MN4 to MN3O or MN3S of metalloporphyrin leads to the formal oxidation state of the central metal to +1. If one wants to maintain a +2 oxidation state of M, an additional axial ligand, typically a Cl atom, is often introduced. Similarly, when two NH groups are replaced to form N2S2 or N2O2-coordinated metalloporphyrins, the central metal is coordinated by two Cl ligands. Experimentally, a few O/S-coordinated metalloporphyrins have been synthesized, including Fe/Co/Ni/Cu/Zn/Mn/Re-21-oxaporphyrin (N3O) [90,91,92,93,94], Li/Fe/Ni/Cu/Ru/Rh/Pd/Re/Hg-21-thiaporphyrin (N3S) [93,95,96,97,98,99,100], Ni-21,23-dioxaporphyrin (N2O2) [90], and Ru/Re-21,23-dithiaporphyrin (N2S2) [101,102]. To date, no experimental studies have reported the successful synthesis of metalloporphyrin complexes incorporating three heteroatoms, such as N1O3 or N1S3. Moreover, MOFs or COFs that incorporate metalloporphyrins with oxygen or sulfur have not yet been reported, although their synthesis looks feasible. In principle, oxygen and sulfur can be interchanged and combined to form metalloporphyrin complexes, such as MN2OS porphyrins. Such structures have not yet been synthesized experimentally, but they have been studied by computational works [67].
Latos-Grażyński and co-workers made pioneering efforts in the field of heteroatom-containing porphyrins synthesis. They reported the first synthesis of Cu(II), Fe(II), and Ni(II) complexes of 21-thiaporphyrin (N3S–TPP) in 1989 [95]. X-ray diffraction showed that the large radius of the S atom disrupts the planar structure in MN4 porphyrins. All these three metalloporphyrins contain an axial chloride ligand, so as to maintain the the metal center in the +II oxidation state. For NiII(–Cl)N3S–TPP, Latos-Grażyński and co-workers carried out chemical and electrochemical reduction to prepare its one-electron reduced NiIN3S-TPP (1989) [103]. The same group synthesized RhIII(–Cl2)N3S–TPP in 1989 [96] and PdII(–Cl)N3S–TPP(–CH3)2 in 1994 [97]. In 2000, LiN3S–TPP(–CH3)2 was synthesized by Arnold et al. [98]. HgII(–Cl)N3S–TPP was prepared by Hwang et al. in 2002 [99]. Later, RuII(–Cl)(–CO)N3S–TPP(–CH3)4 (2011) [100] and ReI(–(CO)3)N3S–TPP (2012) [93] were reported by Hung et al. and Ghosh et al., respectively. For N2S2-porphyrins, the large atomic radius of S atom disrupts the planar stability, so only N2S2-porphyrins containing the heavy metal Ru and Re have been synthesized so far, i.e., RuII(–Cl2)N2S2–TPP(–CH3)4 (2001) [101] and ReI(–(CO)3)N2S2–TPP(–CH3)4 (2014) [102]. These S-substituted metalloporphyrins are illustrated in Figure 7.
The first example of O-coordinated metalloporphyrins was synthesized in 1997. Latos-Grażyński and co-workers synthesized a five-coordinated Ni(II)-21-oxaporphyrin with an axial chloride ligand, Ni(–Cl)N3O–TPP, in which the N3O moiety adopts a planar configuration [90]. Subsequently, they conducted an electroreduction experiment to convert it into a four-coordinated NiIN3O-TPP. In 2002, the same group also reported the formation of FeIII-21-oxaporphyrin [91]. In 2012, Ghosh et al. synthesized ReI(–(CO)3)N3O–TPP [93]. In 2013, Gloe and co-workers reported the Mn(II), Co(II), Zn(II) and Cu(II) complexes of N3O porphyrins [94]. For the N2O2 system, NiII-21,23-dioxaporphyrin was prepared in 1997 [90]. There are two chloride ligands on Ni, and the NiN2O2 motif attains a planar structure. These O-substituted metalloporphyrins are illustrated in Figure 8. The chemical properties of heteroatom-containing porphyrins and metalloporphyrins have been summarized in several review articles [46,47,48,104,105].
Studies concerning the catalytic performances of heteroatom-coordinated metalloporphyrins were only available recently. In 2021, Choi and co-workers investigated the effects of a mono-oxygen substitution in nickel tetraphenylporphyrin (NiN4–TPP) on electrocatalytic CO2 reduction [27]. The modified nickel-21-oxatetraphenylporphyrin, the Ni(–Cl)N3O–TPP catalyst (Figure 9a), was synthesized via a tailored Lindsey method using 2,5-bis (phenylhydroxymethyl) furan as the precursor, followed by metalation and characterization. Electrodes were prepared by mixing the porphyrins with carbon black, and the CO2RR was conducted in a 0.5 M KHCO3 solution. Ni(–Cl)N3O–TPP displays much better electrocatalytic CO2RR activity and selectivity than NiN4–TPP. NiN4–TPP exhibits predominant H2 formation over the whole potential range, but minor CO formation is detected at a high overpotential region below −0.75 VRHE with a maximum FECO of only ca. 2%. (Figure 9c). In sharp contrast, Ni(–Cl)N3O–TPP shows an onset potential of CO formation at approximately −0.55 VRHE, and its maximum FECO reaches ca. 80% at −0.65 VRHE.
The improved performance was attributed to the broken D4h symmetry in Ni(−Cl)N3O−TPP, which results in an increase in the Ni redox potential yielding NiI. NiN4-TPP shows no reduction between −0.4 and −0.8 VSHE, while Ni(–Cl)N3O–TPP is reduced at −0.62 VSHE. This is also supported by DFT calculations, where their respective calculated reduction potential is −1.04 and −0.54 VSHE. As a result, the catalytically active species are different under the applied potential. NiII in NiN4–TPP and NiI in NiN3O–TPP are the actual chemical species responsible for the CO2RR. In DFT calculations using the M06 functional, the authors assumed a CPET mechanism under an applied potential of –0.6 VRHE, and the reaction follows the steps of CO2 adsorption, the formation of *COOH and *CO, and the desorption of CO (Figure 9b). The [NiN3O–TPP] greatly stabilizes the critical species *COOH (Figure 9d), thus leading to a much lower overpotential. In comparison, the calculated COOH binding energy for NiN4-TPP is −0.77 eV, and it is −1.73 eV for [Ni–N3O–TPP]. The authors pointed out that the weak ligand field formed by N3O coordination reduced the energy required for Ni to redistribute electrons to the ligand, thus increasing its Lewis acidity for forming a strong Ni–C bond and stabilizing the *COOH intermediate. At the same time, the lowering of the empty Ni d orbital energy due to this electron redistribution results in a positive shift in its reduction potential. Meanwhile, the author also found that the Ni(–Cl)N3O–TPP complex suffers from instability over a potential of <−0.6 VRHE, posing a challenge for its practical usage. This combined spectroscopic and computational study revealed that the broken ligand-field symmetry is the key for active CO2 electrolysis, and controlling the ligand-field strength would be a challenge for future synthesis.
In 2024, He and coworkers [35] reported similar nickel porphyrin catalyst systems in photocatalytic CO2 reduction. They incorporated O and S into tetracarboxyphenyl ([–TPP–(COOH)4]) nickel porphyrin (NiN4-Por) and synthesized Ni(Cl)ON3-Por and Ni(Cl)SN3-Por. Cyclic voltammetry indicated that the reduction potentials of O/S-substituted species are more positive than the original N4-Por. The values are −1.52 V vs. Fc/Fc+ for NiN4-Por, −1.17 V for Ni(Cl)ON3-Por, and −0.82 V for Ni(Cl)SN3-Por. DFT calculations at the B3LYP-D3 level showed that the lowest unoccupied molecular orbitals (LUMOs) are lowered from NiN4-Por (−2.41 eV) to Ni(Cl)ON3Por (−2.87 eV) and Ni(Cl)SN3Por (−3.06 eV). The lowering of the LUMO suggests higher electron-accepting capability, and it was used to support the positive shift in reduction potential observed in experiment. Although it is possible to give the calculated reduction potential values, the work did not report them.
Then, the authors conducted photocatalytic experiments with a 465 nm LED lamp in CH3CN/H2O mixture (v/v, 4:1) solution and with [Ru(bpy)3]Cl2 as the photosensitizer and triethanolamine (TEOA) as the electron sacrificial agent. Ni(Cl)ON3-Por and Ni(Cl)SN3-Por have remarkable CO2RR activity, with a CO production rate of 24.7 and 38.8 mmol g−1 h−1 with a selectivity of 94.0 and 96.4%, respectively (with H2 as the by-product), outweighing that of NiN4–Por (a CO production rate of 6.6 mmol g−1 h−1 and selectivity of 82.8%). At the same time, the authors also examined the CO production when using H2N4–Por, HON3–Por, and HSN3P–Por as catalysts. They are far less than the CO production from the Ni-porphyrin counterparts, suggesting the Ni center to be the active site. Since the reduction potential of the photosensitizer, [Ru(bpy)3]Cl2 (PS), PS+/PS* is at −1.30 V vs. Fc/Fc+, the active catalytic species were proposed to be NiII in [NiN4–Por]0, NiI in [NiON3–Por]0 (after one-electron reduction), and NiI in [NiSN3Por] (after two-electron reduction) (Figure 10a–c). Then, these active species followed a CPET mechanism, the PDS was the formation of *COOH, and the energy barrier decreased from 1.30 to 0.67 and 0.58 eV for N4 to N3O and N3S systems (Figure 10d). Similarly, the authors attributed such changes to the disruption of D4h symmetry in O/S-substituted Ni-porphyrin. The calculated energy level of Ni dz2 orbital is elevated in the NiON3–Por and NiSN3–Por (where the writer believed the comparisons should be made between the active species), thus enhancing the adsorption of CO2 (Figure 10e). At the same time, the increased electron density on Ni (supposedly, of the active species) stabilizes the *COOH intermediate (Figure 10f–h) and finally leads to faster CO2-to-CO conversion. This work demonstrates how O/S-substitution can alter the electronic states of Ni-porphyrin catalysts and improve their catalytic performance.
To our knowledge, experimental studies on O- or S-substituted metalloporphyrins for CO2 reduction have been limited to Ni metal centers. Studies that involve other metal centers like Co or Fe are theoretical works.
In 2021, Lu et al. reported a DFT study on Fe-N2S2 tetraphenylporphyrin (TPP) and predicted that it can improve CO2 reduction performance to multiple C1 products, including CO, HCOOH, CH3OH, and CH4 [37]. The systems are molecular complexes. For the computations, the VASP 6.1.1 software was used with the PBE-D3 functional and implicit solvent model (water as solvent). The computational hydrogen electrode (CHE) model was used. Fe–N2S2 porphyrin forms a non-coplanar framework with elongated Fe–N bonds (2.14 Å vs. 2.04 Å in Fe–N4). The band gap of the Fe–N2S2 porphyrin (0.34 eV) is much lower than that of the Fe–N4 porphyrin (1.62 eV), due to the raised orbital level of d-orbitals of Fe and p-orbitals of N. The electronic density of states (DOS) reveals additional Fe d orbital contributions in Fe–N2S2, promoting intermediate adsorption in the axial direction. The Fe atom in Fe–N2S2 was found to have a higher electron density than the Fe atom in Fe–N4 porphyrin. Fe–N2S2 exhibits slightly stronger CO2 adsorption (Eads = –0.20 eV) compared to Fe–N4 (Eads = –0.16 eV), attributed to increased electron transfer from Fe to CO2. Note that the adsorption energy difference is about 0.04 eV or 0.9 kcal/mol, which is within the accuracy of DFT calculations. For the CO2-to-CO pathway, the PDS of both Fe–N4 and Fe–N2S2 systems is the desorption of CO, with respective ΔG of 1.45 eV and 1.11 eV (Figure 11a). For the HCOOH pathway, the PDS is the desorption of HCOOH for both FeN4 and FeN2S2 porphyrins and the respective ΔG of 0.70 eV and 0.38 eV (Figure 11b). For the CH3OH pathway, the PDS is CH3OH desorption for the Fe–N4 porphyrin and *CO→*CHO step for the Fe–N2S2 porphyrin, with the respective ΔG of 0.90 eV and 0.40 eV. For the CH4 pathway, the PDS is the *CO→*CHO step for the Fe–N4 porphyrin, and *CH3OH→*CH3 step for FeN2S2 porphyrin, with the respective ΔG of 0.73 eV and 0.56 eV. As a result, the calculated limiting potentials UL for all pathways were lowered in FeN2S2 in comparison to the Fe–N4 case. The respective values are listed in Table 1.
On the Fe–N4 porphyrin, UL follows the sequence HCOOH > CH4 > CH3OH > CO. On the Fe-N2S2 porphyrin, UL follows the sequence HCOOH > CH3OH > CH4 > CO, with the value of −0.38 V for HCOOH and −0.40 V for CH3OH, surpassing most reported catalysts. At the same time, since their values of UL are very close, the HCOOH and CH3OH product channels are in great competition. Compared to conventional Fe–N4, Fe–N2S2 porphyrin shifts product selectivity toward HCOOH and CH3OH, with reduced energy barriers for key steps such as *CO → *CHO. An important drawback of the Fe–N2S2-system is that the hydrogen evolution also becomes easier. The calculated free energy change of H+ + e → *H → H is 0.59 eV on the Fe–N4 porphyrin and 0.26 on the Fe–N2S2 porphyrin. This work underscores the potential of double-heteroatom coordination strategies in designing high-efficiency electrocatalysts.
In 2023, Qi and co-workers [106] reported a theoretical study investigating a series of O- or S-substituted cobalt porphyrin molecular complexes (Figure 12a, where the peripheral ligands are H atoms) for electrochemical CO2RR. The authors performed DFT calculations using Gaussian 09 with a ωB97XD-D3BJ functional and SMD implicit solvent model with water as the solvent. The computational hydrogen electrode model was used. Based on the Co–N4 porphyrin (Co1), they modified the ring structure and designed three structures, including cobalt corrole (Co2), cobalt octahydroporphyrin (Co3), and cobalt 1,5,9,13-tetraazacyclohexadecane (Co4). In terms of the core modification, they introduced O and S heteroatoms and formed six structures, including cobalt 21-oxaporphyrin (Co5, N3O), cobalt 21,23-dioxaporphyrin (Co6, N2O2), 21,22-dioxaporphyrin (Co7, N2O2), cobalt 21-thiaporphyrin (Co8, N3S), cobalt 21,23-dithiaporphyrin (Co9, N2S2), and cobalt 21,22-dithiaporphyrin (Co10, N2S2). Here, we focus on the comparison between Co5–Co10 and Co1. The authors considered different products of CO2 reduction, including CO, CH3OH, and CH4. The product channel of HCOOH and the HER was not discussed in this work.
Similarly, the core-modification approach decreased the band gap as compared to the CoN4 porphyrin. This serves as a hint that the intramolecular electron transfer of these cobalt catalysts would be improved during the CO2RR process. The PDS of CO2-to-CO reduction on Co1 (N4), Co6 (N2O2), Co7 (N2O2), Co9 (N2S2), and Co10 (N2S2) is the formation of *COOH, where their respective ΔG values are 1.27, 2.12, 1.89, 1.17, and 1.41 eV. The PDS of Co5 (N3O) is CO desorption with ΔG of 1.11 eV, indicating a promising catalyst for the further hydrogenation reactions of CO into CH3OH (UL = −1.04 V) and CH4 (UL = −1.04 V) (Figure 12b,c). The PDS of Co8 (N3S) is CO2 adsorption, with a minimum ΔG of 0.58 eV among the studied systems, indicating the best catalytic activity for CO2-to-CO reduction (Figure 12d). Co8 also has the smallest band gaps (3.40 eV), highlighting its strongest reduction capability. The authors concluded that single O/S-substituted cobalt porphyrin exhibits higher activity than double O/S-substituted cobalt porphyrin, and CoN3S is better than CoN3O. The adsorption energies of CO in the designed complexes were analyzed. Change analyses showed that Co5–Co10 transferred more electrons (−0.046 |e| ~ −0.182 |e|) to CO, in comparison to Co1 (0.149 |e|, indicating electron transfer from CO to Co1), hence, a stronger CO binding, and this facilitates the formation of deeper reduction products like CH3OH and CH4. They also claimed that the broken D4h symmetry altered the electron density of the Co porphyrin and improved CO2RR performance.
In 2024, Ren et al. [67] performed periodic DFT calculations with PBE-D3 functional on O/S-substituted metalloporphyrin-based COFs (Por-COFs) for electrocatalytic CO2 reduction. This work designed 15 types of heteroatomic Por-COFs, featuring M–NxOySz (M = Fe, Co, Ni; x + y + z = 4) centers, including single-substituted models: N3O, N3S, and double-substituted models: N2O2, N2S2, and N2OS (Figure 13a). It focused on the electrocatalytic reduction of CO2 to CO and assumed a CPET mechanism. The PDS for all the target M–NxOySz-Por-COFs is the formation of the *COOH intermediate. Figure 13b summarizes the corresponding theoretical limiting potentials UL. In general, changing the coordination from N4 to NxOySz shifts the ULCO2RR towards the positive direction, meaning an easier reduction and thus higher CO2-to-CO catalytic activity. The majority of the designed M–NxOySz-Por-COFs display a distinct preference for CO2RR. Exceptions are Ni–N3O, Co–N3O, Co–N3S, and Co–N2S2 systems, which show a slight preference for HER. Based on calculated UL values, the four CO2-to-CO catalyst candidates ranked at the top are Co–N2O2, Fe–N2OS, Fe–N2O2, and Co–N3O-Por-COFs. The best designed system, Co–N2O2-Por-COF, achieved a limiting potential (ULCO2RR) of −0.58 V, outperforming the parent Co–N4-Por-COF (ULCO2RR = −0.89 V) (Figure 13c). Electronic structure analysis attributed this improvement to enhanced orbital overlap between Co 3d orbitals and *COOH 2p orbitals, alongside increased charge transfer (0.32 e in Co–N2O2 vs. 0.15 e in Co–N4), stabilizing the critical *COOH intermediate (Figure 13d). In comparison, the UL values of the Ni-based systems are more negative than the Fe- and Co-based catalysts. Ren et al. predicted that the UL follows the order of Ni–N4 (−1.52 V) < Ni–N3O (−1.35 V) < Ni–N3S (−1.08 V). Although the Ni–N3S and Ni–N3O systems were not the best, this prediction aligns with the experimental superiority of Ni(Cl)–SN3Por and Ni(Cl)ON3Por observed in the photocatalytic CO2 reduction [35], affirming the accuracy of this computational work. Furthermore, the machine learning model (Gradient Boosting Regression, GBR) further identified central metal mass (m) and metal-ligand bond lengths (dM-X) as pivotal descriptors, underscoring the synergy between metal identity and coordination environment. This work not only predicts high-performance Por-COF candidates but also establishes coordination engineering as a universal strategy for tailoring COF-based electrocatalysts.
The three computational studies mentioned above, although they involve different peripheral ligands, in part share the same first coordination spheres. All calculations assumed a CPET mechanism, and the computational results showed both similarities and differences. The similarities include that, after O/S coordination, the band gap of the metal porphyrin decreases, the electron density at the metal center increases, and the adsorption energies of key intermediates (such as CO2, CO, COOH, and others) change, leading to different catalytic activities for CO2 reduction.
As for the differences, we highlight the following for instance. For the FeN2S2 system, both Lu et al. and Ren et al. predicted a lower UL for CO2-to-CO conversion compared to the FeN4 system (Table 1). However, the former work (employed cluster model) identified CO desorption as the PDS, whereas the latter work (employed periodic model) identified the COOH formation as the PDS. For the Co–NxOy system, Qi et al. (employed cluster model) suggested that the single-substituted Co–N3O-Por performs better than the double-substituted Co–N2O2-Por system, and, in comparison to Co–NxSy systems, S-substitution is better than O-substitution. However, Ren et al.’s calculations (employed periodic model) yielded the opposite conclusion, favoring the double-substituted Co–N2O2-Por-COF system, and O-substitution is better than S-substitution for Co–N4-system. These discrepancies may arise from ligand effects and, of course, the differences in computational models and methods.
Theoretical studies by Ren et al. in 2023 [71] and 2024 [67] employed the same Por-COF model and DFT method, allowing us to compare the calculated CO2-to-CO performance of N-confused and O/S-substituted Co-Por-COF systems. These systems shared the same PDS, i.e., the formation of the *COOH intermediate. As shown in Table 1, for singly substituted Co-Por-COF, the order of calculated UL (thus the order of catalytic performance) is CoN3O (−0.67 V) > CoN3S (−0.74 V) > CoN3C (−0.76 V) > CoN4 (−1.27 V); for double-substitution, the order of UL is CoN2O2 (−0.58 V) > CoN2C2 (−0.60 V) > CoN2S2 (−0.88 V). Based on these UL values, all these systems improved the catalytic performance as compared to the original CoN4 system, because O, S, and C substitutions increased the electron density of the Co center, thereby strengthening the binding with *COOH. CoN2O2 is the best, and double substitution is better than single substitution for C and O, but the trend reverses for S. This latter decline was attributed to the fact that the large atomic radius of S distorts the catalyst structure, reducing orbital overlap and weakening binding to *COOH. This underscores the necessity of balancing electronic modulation and structural stability in heteroatom-substitution strategies, providing crucial guidance for the rational design of high-performance CO2 reduction catalysts.

5. Summary and Outlook

In this review, we have summarized the recent progress in the application of N-confused and O/S-substituted metalloporphyrins-based systems for CO2 reduction, both experimentally and theoretically. In general, the first coordination sphere modification by C, O, and S atoms represents an effective method to improve the catalytic activity and selectivity of CO2 reduction. By breaking the symmetry of traditional M–N4 coordination, these heteroatoms induce distinct electronic perturbations that optimize intermediate adsorption and lower reaction energy barriers. Furthermore, these heteroatom coordination structures significantly influence product selectivity. The catalytic performances of these systems are summarized in Table 2, where properties including reduction potential, products and yields, and Faradaic efficiencies are present. It shows that the reactivity changes depend on the type of substituted atoms, the metal center, and experimental conditions. In addition, computational models and methods also affect the results.
Experimentally, a Cu–NCP complex was found to improve the electroreduction of CO2-to-CH4 products under pulsed potential electrolysis (PPE), mainly because PPE lowered the local pH. A Ni–NCP complex was reported to improve the photoreduction of CO2-to-CO, and it was attributed to a ligand-involved mechanism. In addition, Ni–N3(Cl)O- and/or Ni–N3(Cl)S-porphyrin systems were found to improve the activity and selectivity of CO2-to-CO electrochemically and photochemically. The detailed underlying mechanisms of enhanced catalytic reactivity are different for C and O/S substitution, as well as for Cu and Ni. For the Ni-porphyrin system, common findings include that the C/O/S substitutions shift the reduction potential to positive, the NiI species are the active catalytic species, and S substitution is better than O substitution.
Both electrocatalytic works mentioned that the modified porphyrin systems experienced critical instability issues, indicating that the asymmetric coordination may weaken macrocyclic stability under redox cycling. Under high potential, Cu-NCP electrolysis to Cu clusters/nanoparticles, applying PPE only delayed and cannot avoid the decomplexation of Cu-NCP [89]. The Ni ions were dissolved from the Ni(–Cl)N3O-porphyrin system to form metallic Ni or oxidized NiO and Ni(OH)2 precipitations [27]. In periodic DFT calculations, the stability of original catalysts is often predicted by computing the formation energy and dissolution potential (Udiss) of the metalloporphyrin materials. Predicting the stability of catalysts during the catalytic process, i.e., under applied potential, is normally beyond the calculation capability, due to the difficulty in modeling the real working conditions. To account for the role of solvent in affecting the stability of the catalysts, one may include explicit solvent molecules, but the computational cost is much higher. Moreover, including the effect of potential during the simulations is computationally challenging.
In contrast, both photocatalytic works reported that the catalysts are stable under the experimental conditions. It seems to indicate that, in terms of stability, the first coordination modified porphyrins are suitable for photocatalysis reactions, rather than electrocatalytic reactions.
Computational works extended the systems from single substitution to double and mixed substitutions. Most works focus on Fe, Co, and Ni metal centers. Although the computational models and methods are not the same in those works, the calculation results revealed a number of universal observations, no matter whether the substituted atoms are C, O, or S. The heteroatom substitution changes (mainly increases) the electron density of the center metals, which may enhance the electron transfer from M to other incoming species. It decreases the HOMO-LUMO gap (or band gap) of the porphyrin systems, either by lowering the LUMO or raising the HOMO. The energy level of the dz2 orbital was raised; this can enhance the interaction with incoming species with a proper orbital orientation. These electron redistributions and the change in orbital levels can lead to a shift in the reduction potential and stabilization of key intermediates (such as CO2, COOH, CO, etc.) during CO2RR processes and then change the reaction barrier. For different metals, heteroatom substitution affects the intermediates differently, so the potential determining steps, and thus the selectivity, may differ. The major CO2 reduction product of metalloporphyrin complexes is CO and H2. After introducing heteroatoms in the first coordination sphere, calculations suggest that Ni systems still prefer CO, Fe systems (like Fe–N2S2) may prefer HCOOH and CH3OH, and Co systems (like Co–N3C1, Co–N2C2) have increased preference for CH3OH and CH4.
Theoretical calculations predicted improved CO2RR activity for Fe- and Co-centered porphyrin with C or O or S coordination, and these predictions have not been validated by experiments yet, which suggests a potential avenue for further experimental exploration. We would like to point out that the calculations mentioned above are based on the thermodynamics results as computed from the DFT method by assuming CPET, and the CHE model was usually used. We believe many important issues are worth further exploring. For example, treat the protonation and electron-transfer steps separately [60,61,62,66,107], examine the transition states (as the system may be kinetically controlled) [64], track the flow of electrons to check the oxidation state and clarify the active catalytic species, consider the mechanism that ligands serve as active sites, account for the solvent effect by including explicit solvent molecules, etc. It is worth mentioning that these calculations did not consider factors like catalyst degradation or mass transport in real-world electrolysis. A direct comparison between the overpotential measured in the experiment and the limiting potentials (UL) calculated by DFT is not anticipated. Nevertheless, comparing the calculated parameters, including UL values and activation energies, across different catalysts provides valuable information on the change in catalytic activity and selectivity.
One may be interested in comparing the influence of the first coordination and outer-sphere modification on the catalytic performance of CO2 reduction. Here, we take the comparison between NiN4–TPP, NiN3O-TPP, and Ni–TPP–NItBu (Ni-1,3-ditert-butylimidazolin-2-ylidenamino-tetraphenylporphyrin) for illustration. Experimentally, the unmodified Ni–N4–TPP exhibits extremely low catalytic activity in CO2 electroreduction. Through the first coordination sphere modification, the O-coordinated Ni–N3O–TPP achieved an FECO of 80% at −0.65 VRHE and experienced a positive shift in reduction potential [27]. Through outer-sphere modification, adding the NItBu group to peripheral benzene ring forms Ni–TPP–NItBu. It increased the FECO to 62% at −1.10 VRHE but had no obvious effect on the reduction potential [64]. Both strategies improved the CO2-to-CO reduction performance, but the underlying mechanisms are not the same. For the former, O-substitution disrupts the D4h symmetry of NiN4 and redistributes the electron density on Ni. This promotes the formation of the active species NiI and stabilizes the *COOH intermediate, thus enhancing the CO2 to CO reduction. For the latter, appending the NItBu group on the outer-sphere increases the electron density at the Ni center, for NItBu is a strong electron-donating group. This facilitates the binding of CO2 and improves catalytic performance. We can see that the first coordination sphere affects the electron density of the central metal through ligand-field regulation, whereas the outer-coordination sphere affects it via longer-distance electronic interactions.
In comparison, O-substitution is more effective than appending peripheral NItBu groups in increasing the catalytic reactivity towards CO2-to-CO reduction. However, based on this single comparison, we cannot draw a general conclusion that first coordination modification is more effective than outer-sphere modification. Electron-withdrawing or cationic groups were found to have distinct effects as compared to electron-donating groups. The effects of different substituents were studied for Fe and Co porphyrin systems but not for Ni porphyrin systems [33,34]. We anticipate that combining these two strategies may achieve the maximum enhancement of catalytic performance. In summary, heteroatom-containing metalloporphyrins represent a versatile platform for catalysis for CO2 and other small molecules. This approach not only provides a deeper understanding of the structure and activity relationships governing reduction reactions but also offers a clear pathway for the rational design of next-generation electrocatalysts. We propose the following promising directions for future investigations: the synergistic effect between the modification of first- and peripheral coordination sphere modification [108], the incorporation of multi-metal centers within expanded porphyrin analogs [109,110,111], and the integration with composite materials such as carbon nanostructures [112,113], covalent organic frameworks [114], metal–organic frameworks [115,116], or conductive polymers [117,118]. Furthermore, to establish catalyst design principles, mechanistic investigations require not only theoretical and computational approaches but also the experimental monitoring of dynamic processes. Operando spectroscopy (e.g., XAS, SEIRAS, DEMS) has been applied to study the catalytic mechanism of CO2 reduction [118,119]. These techniques can capture the reaction intermediates and obtain critical information such as reaction rates, product formation kinetics, and the oxidation states of active sites on catalysts, thereby bridging the gap with theoretical calculations and facilitating better catalyst design.

Author Contributions

Conceptualization, J.X.; investigation, Z.L. and Q.W.; writing—original draft preparation, Z.L. and Z.R.; writing—review and editing, J.X. and Z.R.; funding acquisition, J.X. and Z.R. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Key Research and Development Program of China (2023YFC3310600), the Teli Fellowship from Beijing Institute of Technology, Xiaomi Young Scholarship, China, and the Hundred Outstanding Talent Program of Jining University (No. 2024ZYRC109).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing is not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Chu, S.; Majumdar, A. Opportunities and Challenges for a Sustainable Energy Future. Nature 2012, 488, 294–303. [Google Scholar] [CrossRef] [PubMed]
  2. Callahan, C.W. Present and Future Limits to Climate Change Adaptation. Nat. Sustain. 2025, 8, 336–342. [Google Scholar] [CrossRef]
  3. Suri, D.; Das, S.; Choudhary, S.; Venkanna, G.; Sharma, B.; Afroz, M.A.; Tailor, N.K.; Joshi, R.; Satapathi, S.; Tripathi, K. Enigma of Sustainable CO2 Conversion to Renewable Fuels and Chemicals Through Photocatalysis, Electrocatalysis, and Photoelectrocatalysis: Design Strategies and Atomic Level Insights. Small 2025, 21, 2408981. [Google Scholar] [CrossRef] [PubMed]
  4. Qiao, J.; Liu, Y.; Hong, F.; Zhang, J. A Review of Catalysts for the Electroreduction of Carbon Dioxide to Produce Low-Carbon Fuels. Chem. Soc. Rev. 2014, 43, 631–675. [Google Scholar] [CrossRef]
  5. Birdja, Y.Y.; Pérez-Gallent, E.; Figueiredo, M.C.; Göttle, A.J.; Calle-Vallejo, F.; Koper, M.T.M. Advances and Challenges in Understanding the Electrocatalytic Conversion of Carbon Dioxide to Fuels. Nat. Energy 2019, 4, 732–745. [Google Scholar] [CrossRef]
  6. Jin, S.; Hao, Z.; Zhang, K.; Yan, Z.; Chen, J. Advances and Challenges for the Electrochemical Reduction of CO2 to CO: From Fundamentals to Industrialization. Angew. Chem. Int. Ed. 2021, 60, 20627–20648. [Google Scholar] [CrossRef]
  7. She, X.; Wang, Y.; Xu, H.; Chi Edman Tsang, S.; Ping Lau, S. Challenges and Opportunities in Electrocatalytic CO2 Reduction to Chemicals and Fuels. Angew. Chem. Int. Ed. 2022, 61, e202211396. [Google Scholar] [CrossRef]
  8. Grim, R.G.; Huang, Z.; Guarnieri, M.T.; Ferrell, J.R.; Tao, L.; Schaidle, J.A. Transforming the Carbon Economy: Challenges and Opportunities in the Convergence of Low-Cost Electricity and Reductive CO2 Utilization. Energy Environ. Sci. 2020, 13, 472–494. [Google Scholar] [CrossRef]
  9. Xia, J.; Wang, B.; Di, J.; Li, Y.; Yang, S.-Z.; Li, H.; Guo, S. Construction of Single-Atom Catalysts for Electro-, Photo- and Photoelectro-Catalytic Applications: State-of-the-Art, Opportunities, and Challenges. Mater. Today 2022, 53, 217–237. [Google Scholar] [CrossRef]
  10. Ali, S.A.; Sadiq, I.; Estrader, M.; Ahmad, T. Atomic-Level Tuning Strategies in Designing Active Catalysts for Heterogeneous CO2 Conversion into Chemical Feedstock. ChemCatChem 2025, 17, e202500032. [Google Scholar] [CrossRef]
  11. Zhong, S.; Bird, A.; Kopec, R.E. The Metabolism and Potential Bioactivity of Chlorophyll and Metallo-chlorophyll Derivatives in the Gastrointestinal Tract. Mol. Nutr. Food Res. 2021, 65, e2000761. [Google Scholar] [CrossRef] [PubMed]
  12. Osman, D.; Cooke, A.; Young, T.R.; Deery, E.; Robinson, N.J.; Warren, M.J. The Requirement for Cobalt in Vitamin B12: A Paradigm for Protein Metalation. Biochim. Biophys. Acta (BBA)—Mol. Cell Res. 2021, 1868, 118896. [Google Scholar] [CrossRef] [PubMed]
  13. Li, Y.; Wang, N.; Lei, H.; Li, X.; Zheng, H.; Wang, H.; Zhang, W.; Cao, R. Bioinspired N4-Metallomacrocycles for Electrocatalytic Oxygen Reduction Reaction. Coord. Chem. Rev. 2021, 442, 213996. [Google Scholar] [CrossRef]
  14. Zhang, R.; Warren, J.J. Recent Developments in Metalloporphyrin Electrocatalysts for Reduction of Small Molecules: Strategies for Managing Electron and Proton Transfer Reactions. ChemSusChem 2021, 14, 293–302. [Google Scholar] [CrossRef]
  15. Lucas, D.Z.; Isaltino, A.B.; Paulo, C.d.S.F.; Fabricio, B.Z.; Lucas, B.B.; Osvaldo, A.S.; Yassuko, I. Metalloporphyrins in Drug and Pesticide Catalysis as Powerful Tools to Elucidate Biotransformation Mechanisms. Mini-Rev. Org. Chem. 2016, 13, 281–288. [Google Scholar] [CrossRef]
  16. Bonin, J.; Maurin, A.; Robert, M. Molecular Catalysis of the Electrochemical and Photochemical Reduction of CO2 with Fe and Co Metal Based Complexes. Recent Advances. Coord. Chem. Rev. 2017, 334, 184–198. [Google Scholar] [CrossRef]
  17. Zhang, W.; Hu, Y.; Ma, L.; Zhu, G.; Wang, Y.; Xue, X.; Chen, R.; Yang, S.; Jin, Z. Progress and Perspective of Electrocatalytic CO2 Reduction for Renewable Carbonaceous Fuels and Chemicals. Adv. Sci. 2017, 5, 1700275. [Google Scholar] [CrossRef]
  18. Costentin, C.; Passard, G.; Robert, M.; Savéant, J.M. Ultraefficient Homogeneous Catalyst for the CO2-to-CO Electrochemical Conversion. Proc. Natl. Acad. Sci. USA 2014, 111, 14990. [Google Scholar] [CrossRef]
  19. Abdinejad, M.; Wilm, L.F.B.; Dielmann, F.; Kraatz, H.B. Electroreduction of CO2 Catalyzed by Nickel Imidazolin-2-ylidenamino-Porphyrins in Both Heterogeneous and Homogeneous Molecular Systems. ACS Sustain. Chem. Eng. 2020, 9, 521–530. [Google Scholar] [CrossRef]
  20. Lin, S.; Diercks, C.S.; Zhang, Y.-B.; Kornienko, N.; Nichols, E.M.; Zhao, Y.; Paris, A.R.; Kim, D.; Yang, P.; Yaghi, O.M.; et al. Covalent Organic Frameworks Comprising Cobalt Porphyrins for Catalytic CO2 Reduction in Water. Science 2015, 349, 1208–1213. [Google Scholar] [CrossRef]
  21. Kornienko, N.; Zhao, Y.; Kley, C.S.; Zhu, C.; Kim, D.; Lin, S.; Chang, C.J.; Yaghi, O.M.; Yang, P. Metal–Organic Frameworks for Electrocatalytic Reduction of Carbon Dioxide. J. Am. Chem. Soc. 2015, 137, 14129–14135. [Google Scholar] [CrossRef] [PubMed]
  22. Wu, Q.; Xie, R.-K.; Mao, M.-J.; Chai, G.-L.; Yi, J.-D.; Zhao, S.-S.; Huang, Y.-B.; Cao, R. Integration of Strong Electron Transporter Tetrathiafulvalene into Metalloporphyrin-Based Covalent Organic Framework for Highly Efficient Electroreduction of CO2. ACS Energy Lett. 2020, 5, 1005–1012. [Google Scholar] [CrossRef]
  23. Gao, F.; Wu, Y.-P.; Wu, X.-Q.; Li, D.-S.; Yang, G.; Wang, Y.-Y. Transition-Metal Porphyrin-Based MOFs In Situ-Derived Hybrid Catalysts for Electrocatalytic CO2 Reduction. Inorg. Chem. 2024, 63, 8948–8957. [Google Scholar] [CrossRef]
  24. Younis, S.A.; Lim, D.-K.; Kim, K.-H.; Deep, A. Metalloporphyrinic metal-organic frameworks: Controlled synthesis for catalytic applications in environmental and biological media. Adv. Colloid Interface Sci. 2020, 277, 102108. [Google Scholar] [CrossRef] [PubMed]
  25. Xu, N.; Diao, Y.; Qin, X.; Xu, Z.; Ke, H.; Zhu, X. Donor–Acceptor Covalent Organic Frameworks of Nickel(II) Porphyrin for Selective and Efficient CO2 Reduction into CO. Dalton Trans. 2020, 49, 15587–15591. [Google Scholar] [CrossRef]
  26. Domingo-Tafalla, B.; Chatterjee, T.; Palomares, E. Recent Advances in the Rational Designing of Metalloporphyrinoid-Based CO2 Reduction Catalysts: From Molecular Structural Tuning to the Application in Catalysis. J. Porphyr. Phthalocyanines 2023, 27, 23–46. [Google Scholar] [CrossRef]
  27. Kim, H.; Shin, D.; Yang, W.; Won, D.H.; Oh, H.-S.; Chung, M.W.; Jeong, D.; Kim, S.H.; Chae, K.H.; Ryu, J.Y.; et al. Identification of Single-Atom Ni Site Active toward Electrochemical CO2 Conversion to CO. J. Am. Chem. Soc. 2021, 143, 925–933. [Google Scholar] [CrossRef]
  28. Kadish, K.M.; Smith, K.M.; Guilard, R. (Eds.) The Porphyrin Handbook; Academic Press: New York, NY, USA, 2000; Volume 1. [Google Scholar]
  29. Kuznetsov, A.E. Core-modified porphyrins: Novel building blocks in chemistry. Phys. Sci. Rev. 2023, 8, 1513–1543. [Google Scholar] [CrossRef]
  30. Yong, C.; Song, C.; Lu, G.; Wang, X.; Song, G.; Shi, G.; Wang, Y.; Xie, X.; Sun, J. Effect of Axial Ligands on Photocatalytic CO2 Reduction of Zirconium-Based Porphyrins. ACS Appl. Energy Mater. 2025, 8, 4379–4386. [Google Scholar] [CrossRef]
  31. He, H.; Qiu, Z.-Y.; Yin, Z.; Kong, J.; Dang, J.-S.; Lei, H.; Zhang, W.; Cao, R. The meso-substituent electronic effect of Fe porphyrins on the electrocatalytic CO2 reduction reaction. Chem. Commun. 2024, 60, 5916–5919. [Google Scholar] [CrossRef]
  32. Sonea, A.; Warren, J.J. Pendant Catechol Group Improves the Performance of Iron Porphyrin CO2 Reduction Catalysts. ACS Catal. 2025, 15, 1444–1454. [Google Scholar] [CrossRef]
  33. 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, 16639–16644. [Google Scholar] [CrossRef] [PubMed]
  34. Zhu, M.; Yang, D.-T.; Ye, R.; Zeng, J.; Corbin, N.; Manthiram, K. Inductive and electrostatic effects on cobalt porphyrins for heterogeneous electrocatalytic carbon dioxide reduction. Catal. Sci. Technol. 2019, 9, 974–980. [Google Scholar] [CrossRef]
  35. Zhong, Y.H.; Wang, Y.; Zhao, S.Y.; Xie, Z.X.; Chung, L.H.; Liao, W.M.; Yu, L.; Wong, W.Y.; He, J. Regulating the Electronic Configuration of Ni Sites by Breaking Symmetry of Ni-Porphyrin to Facilitate CO2 Photocatalytic Reduction. Adv. Funct. Mater. 2024, 34, 2316199. [Google Scholar] [CrossRef]
  36. Yuan, H.; Krishna, A.; Wei, Z.; Su, Y.; Chen, J.; Hua, W.; Zheng, Z.; Song, D.; Mu, Q.; Pan, W.; et al. Ligand-Bound CO2 as a Nonclassical Route toward Efficient Photocatalytic CO2 Reduction with a Ni N-Confused Porphyrin. J. Am. Chem. Soc. 2024, 146, 10550–10558. [Google Scholar] [CrossRef]
  37. Cao, S.; Wei, S.; Wei, X.; Zhou, S.; Chen, H.; Hu, Y.; Wang, Z.; Liu, S.; Guo, W.; Lu, X. Can N, S Cocoordination Promote Single Atom Catalyst Performance in CO2RR? Fe-N2S2 Porphyrin versus Fe-N4 Porphyrin. Small 2021, 17, e2100949. [Google Scholar] [CrossRef]
  38. Hou, X.; Ding, J.; Liu, W.; Zhang, S.; Luo, J.; Liu, X. Asymmetric Coordination Environment Engineering of Atomic Catalysts for CO2 Reduction. Nanomaterials 2023, 13, 309. [Google Scholar] [CrossRef]
  39. Yang, F.; Han, H.; Duan, H.; Fan, F.; Chen, S.; Yu Xia, B.; He, Y.-L. A Review on Single Site Catalysts for Electrochemical CO2 Reduction. Adv. Energy Mater. 2025, 15, 2405726. [Google Scholar] [CrossRef]
  40. Tang, T.; Wang, Z.; Guan, J. Structural Optimization of Carbon-Based Diatomic Catalysts towards Advanced Electrocatalysis. Coord. Chem. Rev. 2023, 492, 215288. [Google Scholar] [CrossRef]
  41. Verma, S.; Yadav, A. Emerging Single-Atom Catalysts and Nanomaterials for Photoelectrochemical Reduction of Carbon Dioxide to Value-Added Products: A Review of the Current State-of-the-Art and Future Perspectives. Energy Fuels 2023, 37, 5712–5742. [Google Scholar] [CrossRef]
  42. Cui, J.; Yao, Z.; Zhu, B.; Tang, J.; Sun, J.; Xiong, P.; Fu, Y.; Zhang, W.; Zhu, J. Coordination Microenvironment Regulation in Carbon−Based Atomically Dispersed Metal Catalysts for Clean Energy−Conversion. Adv. Funct. Mater. 2024, 34, 2405328. [Google Scholar] [CrossRef]
  43. Manbeck, G.F.; Fujita, E. A Review of Iron and Cobalt Porphyrins, Phthalocyanines and Related Complexes for Electrochemical and Photochemical Reduction of Carbon Dioxide. J. Porphyr. Phthalocyanines 2015, 19, 45–64. [Google Scholar] [CrossRef]
  44. Gotico, P.; Halime, Z.; Aukauloo, A. Recent Advances in Metalloporphyrin-Based Catalyst Design towards Carbon Dioxide Reduction: From Bio-Inspired Second Coordination Sphere Modifications to Hierarchical Architectures. Dalton Trans. 2020, 49, 2381–2396. [Google Scholar] [CrossRef] [PubMed]
  45. Dedic, D.; Dorniak, A.; Rinner, U.; Schofberger, W. Recent Progress in (Photo-)-Electrochemical Conversion of CO2 With Metal Porphyrinoid-Systems. Front. Chem. 2021, 9, 685619. [Google Scholar] [CrossRef]
  46. Chatterjee, T.; Shetti, V.S.; Sharma, R.; Ravikanth, M. Heteroatom-Containing Porphyrin Analogues. Chem. Rev. 2017, 117, 3254. [Google Scholar] [CrossRef]
  47. Gupta, I.; Ravikanth, M. Recent Developments in Heteroporphyrins and Their analogues. Coord. Chem. Rev. 2006, 250, 468–518. [Google Scholar] [CrossRef]
  48. Szyszko, B.; Latos-Grażyński, L. Core chemistry and skeletal rearrangements of porphyrinoids and metalloporphyrinoids. Chem. Soc. Rev. 2015, 44, 3588–3616. [Google Scholar] [CrossRef]
  49. Takahashi, K.; Hiratsuka, K.; Sasaki, H.; Toshima, S. Electrocatalytic Behavior of Metal Porphyrins in the Reduction of Carbon Dioxide. Chem. Lett. 1979, 8, 305–308. [Google Scholar] [CrossRef]
  50. Call, A.; Cibian, M.; Yamamoto, K.; Nakazono, T.; Yamauchi, K.; Sakai, K. Highly Efficient and Selective Photocatalytic CO2 Reduction to CO in Water by a Cobalt Porphyrin Molecular Catalyst. ACS Catal. 2019, 9, 4867–4874. [Google Scholar] [CrossRef]
  51. Yuan, H.; Yu, Y.; Yang, S.; Lei, Q.; Yang, Z.; Lan, B.; Han, Z. Photocatalytic CO2 Reduction with Iron Porphyrin Catalysts and Anthraquinone Dyes. Chem. Commun. 2024, 60, 6292–6295. [Google Scholar] [CrossRef]
  52. Shen, J.; Kortlever, R.; Kas, R.; Birdja, Y.Y.; Diaz-Morales, O.; Kwon, Y.; Ledezma-Yanez, I.; Schouten, K.J.P.; Mul, G.; Koper, M.T.M. Electrocatalytic Reduction of Carbon Dioxide to Carbon Monoxide and Methane at an Immobilized Cobalt Protoporphyrin. Nat. Commun. 2015, 6, 8177. [Google Scholar] [CrossRef]
  53. Amanullah, S.; Saha, P.; Dey, A. Activating the Fe(I) State of Iron Porphyrinoid with Second-Sphere Proton Transfer Residues for Selective Reduction of CO2 to HCOOH via Fe(III/II)–COOH Intermediates. J. Am. Chem. Soc. 2021, 143, 13579–13592. [Google Scholar] [CrossRef] [PubMed]
  54. Liu, Y.; Yang, Y.; Sun, Q.; Wang, Z.; Huang, B.; Dai, Y.; Qin, X.; Zhang, X. Chemical Adsorption Enhanced CO2 Capture and Photoreduction over a Copper Porphyrin Based Metal Organic Framework. ACS Appl. Mater. Interfaces 2013, 5, 7654–7658. [Google Scholar] [CrossRef]
  55. Abdinejad, M.; Seifitokaldani, A.; Dao, C.; Sargent, E.H.; Zhang, X.-a.; Kraatz, H.B. Enhanced Electrochemical Reduction of CO2 Catalyzed by Cobalt and Iron Amino Porphyrin Complexes. ACS Appl. Energy Mater. 2019, 2, 1330–1335. [Google Scholar] [CrossRef]
  56. 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, 9400–9417. [Google Scholar] [CrossRef]
  57. Ifraemov, R.; Mukhopadhyay, S.; Hod, I. Photo-Assisted Electrochemical CO2 Reduction to CH4 Using a Co-Porphyrin-Based Metal–Organic Framework. Sol. RRL 2023, 7, 2201068. [Google Scholar] [CrossRef]
  58. Han, J.; An, P.; Liu, S.; Zhang, X.; Wang, D.; Yuan, Y.; Guo, J.; Qiu, X.; Hou, K.; Shi, L.; et al. Reordering d Orbital Energies of Single-Site Catalysts for CO2 Electroreduction. Angew. Chem. Int. Ed. 2019, 58, 12711–12716. [Google Scholar] [CrossRef] [PubMed]
  59. Fang, J.; Zhu, Y.-N.; Long, X.; Li, X.-B.; Zhang, Q.; Yang, G.; Du, S.; Liu, Z.; Liu, Z.; Peng, F. Theoretical Calculations and Experimental Verification of Carbon Dioxide Reduction Electrocatalyzed by Metalloporphyrin. J. Colloid Interface Sci. 2024, 668, 366–374. [Google Scholar] [CrossRef]
  60. Nielsen, I.M.B.; Leung, K. Cobalt−Porphyrin Catalyzed Electrochemical Reduction of Carbon Dioxide in Water. 1. A Density Functional Study of Intermediates. J. Phys. Chem. A 2010, 114, 10166–10173. [Google Scholar] [CrossRef]
  61. Shen, J.; Kolb, M.J.; Göttle, A.J.; Koper, M.T.M. DFT Study on the Mechanism of the Electrochemical Reduction of CO2 Catalyzed by Cobalt Porphyrins. J. Phys. Chem. C 2016, 120, 15714–15721. [Google Scholar] [CrossRef]
  62. Miyamoto, K.; Asahi, R. Water Facilitated Electrochemical Reduction of CO2 on Cobalt-Porphyrin Catalysts. J. Phys. Chem. C 2019, 123, 9944–9948. [Google Scholar] [CrossRef]
  63. Bochlin, Y.; Ben-Eliyahu, Y.; Kadosh, Y.; Kozuch, S.; Zilbermann, I.; Korin, E.; Bettelheim, A. DFT and Empirical Considerations on Electrocatalytic Water/Carbon Dioxide Reduction by CoTMPyP in Neutral Aqueous Solutions**. ChemPhysChem 2020, 21, 2644–2650. [Google Scholar] [CrossRef]
  64. Zhang, Y.-Q.; Chen, J.-Y.; Siegbahn, P.E.M.; Liao, R.-Z. Harnessing Noninnocent Porphyrin Ligand to Circumvent Fe-Hydride Formation in the Selective Fe-Catalyzed CO2 Reduction in Aqueous Solution. ACS Catal. 2020, 10, 6332–6345. [Google Scholar] [CrossRef]
  65. Cao, Y.C.; Shi, L.L.; Li, M.; You, B.; Liao, R.Z. Deciphering the Selectivity of the Electrochemical CO2 Reduction to CO by a Cobalt Porphyrin Catalyst in Neutral Aqueous Solution: Insights from DFT Calculations. ChemistryOpen 2023, 12, e202200254. [Google Scholar] [CrossRef] [PubMed]
  66. Göttle, A.J.; Koper, M.T.M. Proton-Coupled Electron Transfer in the Electrocatalysis of CO2 Reduction: Prediction of Sequential vs. Concerted Pathways using DFT. Chem. Sci. 2017, 8, 458–465. [Google Scholar] [CrossRef]
  67. Ren, Z.; Gong, K.; Zhao, B.; Chen, S.-L.; Xie, J. Boosting the Catalytic Performance of Metalloporphyrin-based Covalent Organic Frameworks via Coordination Engineering for CO2 and O2 Reduction. Mater. Chem. Front. 2024, 8, 1958–1970. [Google Scholar] [CrossRef]
  68. Wang, Y.-R.; Huang, Q.; He, C.-T.; Chen, Y.; Liu, J.; Shen, F.-C.; Lan, Y.-Q. Oriented Eectron Transmission in Polyoxometalate-Metalloporphyrin Organic Framework for Highly Selective Electroreduction of CO2. Nat. Commun. 2018, 9, 4466. [Google Scholar] [CrossRef]
  69. Nørskov, J.K.; Rossmeisl, J.; Logadottir, A.; Lindqvist, L.; Kitchin, J.R.; Bligaard, T.; Jónsson, H. Origin of the Overpotential for Oxygen Reduction at a Fuel-Cell Cathode. J. Phys. Chem. B 2004, 108, 17886–17892. [Google Scholar] [CrossRef]
  70. Warburton, R.E.; Soudackov, A.V.; Hammes-Schiffer, S. Theoretical Modeling of Electrochemical Proton-Coupled Electron Transfer. Chem. Rev. 2022, 122, 10599–10650. [Google Scholar] [CrossRef]
  71. Ren, Z.; Zhao, B.; Xie, J. Designing N-Confused Metalloporphyrin-Based Covalent Organic Frameworks for Enhanced Electrocatalytic Carbon Dioxide Reduction. Small 2023, 19, e2301818. [Google Scholar] [CrossRef]
  72. Furuta, H.; Asano, T.; Ogawa, T. “N-Confused Porphyrin”: A New Isomer of Tetraphenylporphyrin. J. Am. Chem. Soc. 1994, 116, 767–768. [Google Scholar] [CrossRef]
  73. Chmielewski, P.J.; Latos-Grażyński, L. N-Methyltetraphenylporphyrin with an Inverted N-Methylpyrrole Ring: The First Isomer of N-Methyltetraphenylporphyrin. J. Chem. Soc. Perkin Trans. II Phys. Org. Chem. 1995, 24, 503–509. [Google Scholar] [CrossRef]
  74. Chmielewski, P.J.; Latos-Grażyński, L.; Schmidt, I. Copper(II) Complexes of Inverted Porphyrin and Its Methylated Derivatives. Inorg. Chem. 2000, 39, 5475. [Google Scholar] [CrossRef] [PubMed]
  75. Chen, W.C.; Hung, C.H. Synthesis and Characterization of Iron N-Confused Porphyrins: Structural Evidences of Agostic Interaction. Inorg. Chem. 2001, 40, 5070–5071. [Google Scholar] [CrossRef]
  76. Bohle, D.S.; Chen, W.C.; Hung, C.H. Metal Oxidation Promoted C-H Activation in Manganese Complexes of N-Confused Porphyrin. Inorg. Chem. 2002, 41, 3334. [Google Scholar] [CrossRef]
  77. Maeda, H.; Osuka, A.; Ishikawa, Y.; Aritome, I.; Hisaeda, Y.; Furuta, H. N-Confused Porphyrin Bearing meso-Perfluorophenyl Groups: A Potential Agent that Forms Stable Square Planar Complexes with Cu(II) and Ag(III). Org. Lett. 2003, 5, 1293. [Google Scholar] [CrossRef]
  78. Zilbermann, I.; Maimon, E.; Ydgar, R.; Shames, A.I.; Korin, E.; Soifer, L.; Bettelheim, A. Spectroscopic and Electrochemical Characterization of Solutions and Films of a New Redox Couple: Co(II)/Co(III) N-Confused Porphyrin. Inorg. Chem. Commun. 2004, 7, 1238. [Google Scholar] [CrossRef]
  79. Harvey, J.D.; Shaw, J.L.; Herrick, R.S.; Ziegler, C.J. The Synthesis of Isostructural Mo2+ Porphyrin and N-Confused Porphyrin Complexes. Chem. Commun. 2005, 10, 4663. [Google Scholar] [CrossRef]
  80. Srinivasan, A.; Toganoh, M.; Niino, T.; Osuka, A.; Furuta, H. Synthesis of N-Confused Tetraphenylporphyrin Rhodium Complexes Having Versatile Metal Oxidation States. Inorg. Chem. 2008, 47, 11305. [Google Scholar] [CrossRef]
  81. Toganoh, M.; Ikeda, S.; Furuta, H. Synthesis, Reactivity, and Properties of N-Fused Porphyrin Rhenium(I) Tricarbonyl Complexes. Inorg. Chem. 2007, 46, 10003. [Google Scholar] [CrossRef]
  82. Furuta, H.; Kubo, N.; Maeda, H.; Ishizuka, T.; Osuka, A.; Nanami, H.; Ogawa, T. N-Confused Double-Decker Porphyrins. Inorg. Chem. 2000, 39, 5424. [Google Scholar] [CrossRef] [PubMed]
  83. Toganoh, M.; Niino, T.; Furuta, H. Luminescent Au(III) Organometallic Complex of N-Confused Tetraphenylporphyrin. Chem. Commun. 2008, 13, 4070. [Google Scholar] [CrossRef]
  84. Peng, C.C.; Yang, F.-A.; Chen, J.-H.; Wang, S.-S.; Tung, J.-Y. Mercury(II) Complex of Inverted N-Methylated Porphyrin: HgPh(2-NCH3NCTPP). Polyhedron 2008, 27, 2309–2314. [Google Scholar] [CrossRef]
  85. Shao, P.; Ren, Z.; Zhao, B.; Wang, X.; Li, J.; Xie, J.; Wang, B.; Feng, X. Theory-Guided Design of N-Confused Porphyrinic Covalent Organic Frameworks for Oxygen Reduction Reaction. J. Am. Chem. Soc. 2025, 147, 8769–8777. [Google Scholar] [CrossRef]
  86. Furuta, H.; Maeda, H.; Osuka, A. Doubly N-Confused Porphyrin: A New Complexing Agent Capable of Stabilizing Higher Oxidation States. J. Am. Chem. Soc. 2000, 122, 803–807. [Google Scholar] [CrossRef]
  87. Maeda, H.; Osuka, A.; Furuta, H. Trans Doubly N-Confused Porphyrins: Cu(III) Complexation and Formation of Rodlike Hydrogen-Bonding Networks. J. Am. Chem. Soc. 2003, 125, 15690–15691. [Google Scholar] [CrossRef] [PubMed]
  88. Huang, S.; Tranca, D.; Rodriguez-Hernandez, F.; Zhang, J.; Lu, C.; Zhu, J.; Liang, H.W.; Zhuang, X. Well-defined N3C1 -anchored Single-Metal-Sites for Oxygen Reduction Reaction. Angew. Chem. Int. Ed. 2024, 63, e202314833. [Google Scholar] [CrossRef]
  89. Hua, W.; Liu, T.; Zheng, Z.; Yuan, H.; Xiao, L.; Feng, K.; Hui, J.; Deng, Z.; Ma, M.; Cheng, J.; et al. Pulse Electrolysis Turns on CO2 Methanation through N-Confused Cupric Porphyrin. Angew. Chem. Int. Ed. 2024, 63, e202315922. [Google Scholar] [CrossRef]
  90. Chmielewski, P.J.; Latos-Grazynski, L.; Olmstead, M.M.; Balch, A.L. Nickel Complexes of 21-Oxaporphyrin and 21, 23-Dioxaporphyrin. Chemistry 1997, 3, 268–278. [Google Scholar] [CrossRef]
  91. Pawlicki, M.; Latos-Grażyński, L. Iron Complexes of 5,10,15,20-Tetraphenyl-21-Oxaporphyrin. Inorg. Chem. 2002, 41, 5866–5873. [Google Scholar] [CrossRef]
  92. Pawlicki, M.; Latos-Grażyński, L. Reactivity of Iron(II) 5,10,15,20-Tetraaryl-21-Oxaporphyrin with Arylmagnesium Bromide:  Formation of Paramagnetic Six-Coordinate Complexes with Two Axial Aryl Groups. Inorg. Chem. 2004, 43, 5564–5571. [Google Scholar] [CrossRef] [PubMed]
  93. Ghosh, A.; Ravikanth, M. Rhenium(I) Tricarbonyl Complexes of 5,10,15,20-Tetraphenyl-21-Thia and 21-Oxaporphyrins. Inorg. Chem. 2012, 51, 6700–6709. [Google Scholar] [CrossRef] [PubMed]
  94. Stute, S.; Götzke, L.; Meyer, D.; Merroun, M.L.; Rapta, P.; Kataeva, O.; Seichter, W.; Gloe, K.; Dunsch, L.; Gloe, K. Molecular Structure, UV/Vis Spectra, and Cyclic Voltammograms of Mn(II), Co(II), and Zn(II) 5,10,15,20-Tetraphenyl-21-oxaporphyrins. Inorg. Chem. 2013, 52, 1515–1524. [Google Scholar] [CrossRef]
  95. Latos-Grazynski, L.; Lisowski, J.; Olmstead, M.M.; Balch, A.L. Five-coordinate Complexes of 21-Thiaporphyrin. Preparations, Spectra, and Structures of Iron(II), Nickel(II), and Copper(II) Complexes. Inorg. Chem. 1989, 28, 1183–1188. [Google Scholar] [CrossRef]
  96. Latos-Grazynski, L.; Lisowski, J.; Olmstead, M.M.; Balch, A.L. Preparation and Structural Characterization of a Six-Coordinate 21-Thiaporphyrin Complex: RhIII(STPP)Cl2 (STPP = tetraphenyl-21-thiaporphyrin anion). Inorg. Chem. 1989, 28, 3328–3331. [Google Scholar] [CrossRef]
  97. Latos-Grazynski, L.; Lisowski, J.; Chmielewski, P.; Grzeszczuk, M.; Olmstead, M.M.; Balch, A.L. Palladium Complexes of 21-Thiaporphyrin: Syntheses and Characterization. Inorg. Chem. 1994, 33, 192–197. [Google Scholar] [CrossRef]
  98. Gebauer, A.; Schmidt, J.A.R.; Arnold, J. Synthesis, Characterization, and Properties of a Lithium 21-Thiaporphyrin Complex. Inorg. Chem. 2000, 39, 3424–3427. [Google Scholar] [CrossRef]
  99. Tung, J.-Y.; Liau, B.-C.; Elango, S.; Chen, J.-H.; Hsieh, H.-Y.; Liao, F.-L.; Wang, S.-L.; Hwang, L.-P. Preparation and Structural Characterization of Mercury 21-Thiaporphyrin Complex: HgII(STPP)Cl (STPP=tetraphenyl-21-thiaporphyrin anion). Inorg. Chem. Commun. 2002, 5, 150–155. [Google Scholar] [CrossRef]
  100. Chuang, C.-H.; Ou, C.-K.; Liu, S.-T.; Kumar, A.; Ching, W.-M.; Chiang, P.-C.; dela Rosa, M.A.C.; Hung, C.-H. Ruthenium Complexes of Thiaporphyrin and Dithiaporphyrin. Inorg. Chem. 2011, 50, 11947–11957. [Google Scholar] [CrossRef]
  101. Hung, C.-H.; Ou, C.-K.; Lee, G.-H.; Peng, S.-M. Structure and Characterization of the First Metal Complex of Dithiaporphyrin:  Ru(S2TTP)Cl2. Inorg. Chem. 2001, 40, 6845–6847. [Google Scholar] [CrossRef]
  102. Kaur, T.; Ghosh, A.; Rajakannu, P.; Ravikanth, M. Synthesis and Crystal Structure of the Rhenium(I) Tricarbonyl Complex of 5,10,15,20-Tetra-p-tolyl-21,23-dithiaporphyrin. Inorg. Chem. 2014, 53, 2355–2357. [Google Scholar] [CrossRef] [PubMed]
  103. Chmielewski, P.; Grzeszczuk, M.; Latos-Grazynski, L.; Lisowski, J. Studies of the Reduction of the Nickel(II) Complex of 5,10,15,20-Tetraphenyl-21-Thiaporphyrin to Form Corresponding Nickel(I) Complexes. Inorg. Chem. 1989, 28, 3546–3552. [Google Scholar] [CrossRef]
  104. Chmielewski, P.J.; Latos-Grażyński, L. Core Modified Porphyrins—A Macrocyclic Platform for Organometallic Chemistry. Coord. Chem. Rev. 2005, 249, 2510–2533. [Google Scholar] [CrossRef]
  105. Pareek, Y.; Ravikanth, M. Thiaporphyrins: From Building Blocks to Multiporphyrin Arrays. RSC Adv. 2014, 4, 7851–7880. [Google Scholar] [CrossRef]
  106. Ding, X.; Jin, Y.; Wang, H.; Qi, D. Molecular Modification of Planar Four-Coordinated Cobalt Active Site for the Electrochemical Reduction of Carbon Dioxide: A Density Functional Theory Study. Inorg. Chem. Front. 2023, 10, 7054–7063. [Google Scholar] [CrossRef]
  107. Göttle, A.J.; Koper, M.T.M. Determinant Role of Electrogenerated Reactive Nucleophilic Species on Selectivity during Reduction of CO2 Catalyzed by Metalloporphyrins. J. Am. Chem. Soc. 2018, 140, 4826–4834. [Google Scholar] [CrossRef]
  108. Costentin, C.; Savéant, J.-M. Towards an Intelligent Design of Molecular Electrocatalysts. Nat. Rev. Chem. 2017, 1, 0087. [Google Scholar] [CrossRef]
  109. Han, J.; Li, N.; Chen, D.; Xu, Q.; Lu, J. Boosting Photocatalytic Activity for Porphyrin-based D-A Conjugated Polymers via Dual Metallic Sites Regulation. Appl. Catal. B Environ. 2022, 317, 121724. [Google Scholar] [CrossRef]
  110. Toganoh, M.; Furuta, H. Creation from Confusion and Fusion in the Porphyrin World—The Last Three Decades of N-Confused Porphyrinoid Chemistry. Chem. Rev. 2022, 122, 8313–8437. [Google Scholar] [CrossRef]
  111. Dong, H.; Fang, L.; Chen, K.-X.; Wei, J.-X.; Li, J.-X.; Qiao, X.; Wang, Y.; Zhang, F.-M.; Lan, Y.-Q. Dual Metallosalen-Based Covalent Organic Frameworks for Artificial Photosynthetic Diluted CO2 Reduction. Angew. Chem. Int. Ed. 2025, 64, e202414287. [Google Scholar] [CrossRef]
  112. Abdinejad, M.; Tang, K.; Dao, C.; Saedy, S.; Burdyny, T. Immobilization Strategies for Porphyrin-Based Molecular Catalysts for the Electroreduction of CO2. J. Mater. Chem. A 2022, 10, 7626–7636. [Google Scholar] [CrossRef] [PubMed]
  113. Gu, S.; Lu, J.; Ren, X.; Wu, N.; Wang, D.; Fang, J.; Zhong, J. Nitrogen-Rich Carbon Nanomaterials Embedded with Ni Nanoparticles for Electrochemical Carbon Dioxide Reduction. ACS Appl. Nano Mater. 2025, 8, 4649–4657. [Google Scholar] [CrossRef]
  114. Cheng, K.; Kong, S.; Wang, J.; Wang, Q.; Yuan, S.; Li, P.-Z.; Zhao, Y. Integrating Multifunctionalities into a 3D Covalent Organic Framework for Efficient CO2 Photoreduction. Angew. Chem. Int. Ed. 2025, 64, e202504772. [Google Scholar] [CrossRef]
  115. Fan, S.-C.; Yuan, W.; Wang, J.-W.; Xing, C.; Li, D.; Wang, Y.; Zhai, Q.-G. Metallic Engineering in Metal-Porphyrin Frameworks for Highly Efficient CO2 Photoreduction. Chem. Eng. J. 2024, 501, 157716. [Google Scholar] [CrossRef]
  116. Han, H.; Cui, H.; Wang, J.; Ge, K.; Yang, Y. Photocatalytic Reduction of CO2 by Multimetallic 2D MOF-Co/Cu/Fe. ChemistrySelect 2024, 9, e202403927. [Google Scholar] [CrossRef]
  117. Zaar, F.; Emanuelsson, R.; Gaiser, P.; Strømme, M.; Sjödin, M. Characterization and Catalytic Prospects of Metalloporphyrin-Functionalized Conducting Polymers. Electrochim. Acta 2023, 467, 143003. [Google Scholar] [CrossRef]
  118. Zhang, H.; Gu, C.; Fan, M.; Zhao, Z.; Kong, X.; Geng, Z. Applications of In-Situ Spectroscopic Techniques towards CO2 Electroreduction. Sci. China Chem. 2024, 67, 3925–3933. [Google Scholar] [CrossRef]
  119. Salamé, A.; Hon Cheah, M.; Bonin, J.; Robert, M.; Anxolabéhère-Mallart, E. Operando Spectroelectrochemistry Unravels the Mechanism of CO2 Electrocatalytic Reduction by an Fe Porphyrin. Angew. Chem. Int. Ed. 2024, 63, e202412417. [Google Scholar] [CrossRef]
Figure 1. Representative structures of (a) porphyrin, (b) metalloporphyrin complexes, (c) N-confused metalloporphyrin complexes and (d) heteroatom-containing metalloporphyrin complexes.
Figure 1. Representative structures of (a) porphyrin, (b) metalloporphyrin complexes, (c) N-confused metalloporphyrin complexes and (d) heteroatom-containing metalloporphyrin complexes.
Molecules 30 02287 g001
Figure 2. (a) Possible CO2RR reaction pathways (under CPET assumption) and products with (b) corresponding overall reaction equations and standard redox potentials [71].
Figure 2. (a) Possible CO2RR reaction pathways (under CPET assumption) and products with (b) corresponding overall reaction equations and standard redox potentials [71].
Molecules 30 02287 g002
Figure 3. Synthesized N-confused metalloporphyrins to date [73,74,75,76,77,78,83,86,87].
Figure 3. Synthesized N-confused metalloporphyrins to date [73,74,75,76,77,78,83,86,87].
Molecules 30 02287 g003
Figure 4. Structural scheme of (a) MN4-Por, MN3C1-Por, MN2C2-Por, and (b) Por-COFs. Free energy diagrams of CO2RR on (c) CoN4-Por-COF, (d) CoN3C1-Por-COF, and (e) CoN2C2-Por-COF. Adapted from reference [71].
Figure 4. Structural scheme of (a) MN4-Por, MN3C1-Por, MN2C2-Por, and (b) Por-COFs. Free energy diagrams of CO2RR on (c) CoN4-Por-COF, (d) CoN3C1-Por-COF, and (e) CoN2C2-Por-COF. Adapted from reference [71].
Molecules 30 02287 g004
Figure 5. Electrocatalytic CO2RR performances of CuTPP and CuNCP under CPE: (a) LSVs of CuNCP and CuTPP taken in the flow cell containing 1M KOH purged with Ar or CO2. CO2RR product distribution and the corresponding FEs at different potentials for (b) CuTPP and (c) CuNCP. (d) Partial current densities of the reduction products at different potentials for CuNCP under CPE [89].
Figure 5. Electrocatalytic CO2RR performances of CuTPP and CuNCP under CPE: (a) LSVs of CuNCP and CuTPP taken in the flow cell containing 1M KOH purged with Ar or CO2. CO2RR product distribution and the corresponding FEs at different potentials for (b) CuTPP and (c) CuNCP. (d) Partial current densities of the reduction products at different potentials for CuNCP under CPE [89].
Molecules 30 02287 g005
Figure 6. (a) Structure of NiNCP. (b) CO evolution amount comparison for NiNCP and NiTPP, FECO for NiNCP. (c) Energy level alignment and electron transfer among the [Ru(bpy)3]2+ photosensitizer, NiTPP and NiNCP cocatalysts, and the CO2/CO redox pair. (d) Free energy profiles for the CO2 reductions with three catalysts. (e) Proposed mechanism of the photocatalytic CO2RR with NiNCP. Adapted from reference [36].
Figure 6. (a) Structure of NiNCP. (b) CO evolution amount comparison for NiNCP and NiTPP, FECO for NiNCP. (c) Energy level alignment and electron transfer among the [Ru(bpy)3]2+ photosensitizer, NiTPP and NiNCP cocatalysts, and the CO2/CO redox pair. (d) Free energy profiles for the CO2 reductions with three catalysts. (e) Proposed mechanism of the photocatalytic CO2RR with NiNCP. Adapted from reference [36].
Molecules 30 02287 g006
Figure 7. Synthesized S-substituted metalloporphyrins to date [93,95,96,97,98,99,100,101,102].
Figure 7. Synthesized S-substituted metalloporphyrins to date [93,95,96,97,98,99,100,101,102].
Molecules 30 02287 g007
Figure 8. Synthesized O-substituted metalloporphyrins to date [90,91,93,94].
Figure 8. Synthesized O-substituted metalloporphyrins to date [90,91,93,94].
Molecules 30 02287 g008
Figure 9. (a) Structures of NiN4-TPP and NiN3O–TPP. (b) Suggested CO2RR mechanism on NiN4–TPP and NiN3O–TPP. (c) FETotal and FECO of CO2-to-CO conversion in an H-type electrochemical cell. (d) DFT-calculated free energy profiles when a potential of −0.6 V vs. RHE is applied. Adapted from reference [27].
Figure 9. (a) Structures of NiN4-TPP and NiN3O–TPP. (b) Suggested CO2RR mechanism on NiN4–TPP and NiN3O–TPP. (c) FETotal and FECO of CO2-to-CO conversion in an H-type electrochemical cell. (d) DFT-calculated free energy profiles when a potential of −0.6 V vs. RHE is applied. Adapted from reference [27].
Molecules 30 02287 g009
Figure 10. (a–c) Schematic illustration of catalytic cycles of NiN4–Por, Ni(Cl)ON3–Por, and Ni(Cl)SN3–Por for CO2-to-CO conversion; (d) the calculated Gibbs free energy diagrams of CO2-to-CO conversion; (e) the calculated adsorption energy of *COOH; (fh) electron density difference plots of NiN4Por, Ni(Cl)ON3Por, and Ni(Cl)SN3Por in rate-determining step (sky blue and orange indicate electron depletion and accumulation, and the isovalue is 0.01) [35].
Figure 10. (a–c) Schematic illustration of catalytic cycles of NiN4–Por, Ni(Cl)ON3–Por, and Ni(Cl)SN3–Por for CO2-to-CO conversion; (d) the calculated Gibbs free energy diagrams of CO2-to-CO conversion; (e) the calculated adsorption energy of *COOH; (fh) electron density difference plots of NiN4Por, Ni(Cl)ON3Por, and Ni(Cl)SN3Por in rate-determining step (sky blue and orange indicate electron depletion and accumulation, and the isovalue is 0.01) [35].
Molecules 30 02287 g010
Figure 11. Free energy diagrams of CO2RR to (a) CO and (b) HCOOH and corresponding intermediates on the Fe–N4 porphyrin and Fe–N2S2 porphyrin (at 0 V vs. RHE). The values in the figures refer to the free energy change for the potential-determining step. The asterisks mean intermediates bind at the active site [37].
Figure 11. Free energy diagrams of CO2RR to (a) CO and (b) HCOOH and corresponding intermediates on the Fe–N4 porphyrin and Fe–N2S2 porphyrin (at 0 V vs. RHE). The values in the figures refer to the free energy change for the potential-determining step. The asterisks mean intermediates bind at the active site [37].
Molecules 30 02287 g011
Figure 12. (a) Structures of Co–X (X = N, O, S). Free energy diagrams of CO2 reduction pathway to (b) CH3OH and (c) CH4 on Co5. (d) Free energy diagrams of CO2RR to CO on Co-X complexes. Adapted from reference [106].
Figure 12. (a) Structures of Co–X (X = N, O, S). Free energy diagrams of CO2 reduction pathway to (b) CH3OH and (c) CH4 on Co5. (d) Free energy diagrams of CO2RR to CO on Co-X complexes. Adapted from reference [106].
Molecules 30 02287 g012
Figure 13. (a) Unit cell of the M–N4-Por-COF (M = Fe/Co/Ni) and the constructed NxOySz coordination models. (b) Limiting potential of CO2RR-to-CO of Fe/Co/Ni–NxOySz-Por-COFs. (c) Free energy diagrams of CO2RR on Co–N4-and Co–N2O2-Por-COFs. (d) Projected electronic densities of states (PDOS) and charge density difference (CDD) of *COOH intermediates adsorbed on Co–N4 and Co–N2O2-Por-COFs. Adapted from reference [67].
Figure 13. (a) Unit cell of the M–N4-Por-COF (M = Fe/Co/Ni) and the constructed NxOySz coordination models. (b) Limiting potential of CO2RR-to-CO of Fe/Co/Ni–NxOySz-Por-COFs. (c) Free energy diagrams of CO2RR on Co–N4-and Co–N2O2-Por-COFs. (d) Projected electronic densities of states (PDOS) and charge density difference (CDD) of *COOH intermediates adsorbed on Co–N4 and Co–N2O2-Por-COFs. Adapted from reference [67].
Molecules 30 02287 g013
Table 1. Calculated limiting potential (UL/V) of the products for the H2, CO, HCOOH, CH3OH and CH4 production on metalloporphyrin systems and DFT methods.
Table 1. Calculated limiting potential (UL/V) of the products for the H2, CO, HCOOH, CH3OH and CH4 production on metalloporphyrin systems and DFT methods.
H2COHCOOHCH3OHCH4DFT MethodRef.
FeN4-TPP-Por−0.59−1.45−0.70−0.90−0.73Cluster
PBE-D3
[37]
FeN2S2-TPP-Por−0.26−1.11−0.38−0.40−0.56
FeN4-Por-COF−1.25−1.25///
FeN3S-Por-COF−1.25−1.18///Periodic
PBE-D3
FeN2S2-Por-COF−1.03−0.91///[67]
FeN3O-Por-COF−1.50−1.24///
FeN2O2-Por-COF−1.26−0.66///
CoN4-Por/−1.27/−0.80−0.80 [106]
CoN3S-Por/−0.58/−0.91−0.91Cluster
ωB97XD-D3BJ
CoN2S2-Por/−1.17/−0.66−1.20
CoN3O-Por/−1.11/−1.04−1.04
CoN2O2-Por/−2.12/−0.96−1.40
CoN4-Por-COF−0.91−0.89−1.36−1.31−1.63 [71]
CoN3C1-Por-COF−0.79−0.76/−0.76−0.99Periodic
PBE-D3
CoN2C2-Por-COF−0.63−0.60/−0.60−1.44
CoN3S-Por-COF−0.75−0.74///Periodic
PBE-D3
CoN2S2-Por-COF−0.84−0.88///[67]
CoN3O-Por-COF−0.64−0.67///
CoN2O2-Por-COF−0.79−0.58///
NiN4-Por-COF−1.52−1.52///
NiN3S-Por-COF−1.16−1.08///
NiN2S2-Por-COF−1.28−1.25///Periodic
PBE-D3
[67]
NiN3O-Por-COF−0.97−1.35///
NiN2O2-Por-COF−1.23−0.98///
Table 2. Summary of different catalyst systems and catalytic performance of metalloporphyrin complexes in CO2RR (Ered: reduction potential; UL: limiting potential; EC: electrocatalysis; PC: photocatalysis; Expt.: experimental; TPP: tetraphenylporphyrin; and Por: porphyrin).
Table 2. Summary of different catalyst systems and catalytic performance of metalloporphyrin complexes in CO2RR (Ered: reduction potential; UL: limiting potential; EC: electrocatalysis; PC: photocatalysis; Expt.: experimental; TPP: tetraphenylporphyrin; and Por: porphyrin).
CatalystsCoresProduct, FEPotentialMethodCat. TypeSolventRef.
Iron-TPPFeN4HCOOH−0.70 VRHE (UL)DFTECH2O[37]
Iron-21,23-dithia-TPPFeN2S2HCOOH
CH3OH
−0.38 VRHE (UL)
−0.40 VRHE (UL)
Cobalt-TPPCoN4CO, 95%−0.60 VRHE (Ered)Expt.EC0.5 M NaHCO3[34]
Cobalt 21-thia-PorCoN3SCO−0.58 VRHE (UL)DFTECH2O[106]
Cobalt 21,23-dithia-PorCoN2S2CH3OH−0.66 VRHE (UL)
Cobalt 21-oxa-PorCoN3OCH3OH, CH4−1.04 VRHE (UL)
Cobalt 21,23-dioxa-PorCoN2O2CH3OH−0.96 VRHE (UL)
CoN4-Por-COFsCoN4CO−0.89 VRHE (UL)DFTECH2O[71]
CoN3C-Por-COFsCoN3CCO, CH3OH−0.76 VRHE (UL)
CoN2C2-Por-COFsCoN2C2CO, CH3OH−0.60 VRHE (UL)
Nickel-TPPNiN4CO, 2%
CO, 29%
−0.75 VRHE (Ered)
−1.10 VRHE (Ered)
Expt.+DFT
Expt.
EC0.5 M KHCO3
DMF
[27]
[19]
Nickel-21-oxa-TPPNiN3OCO, 80%−0.65 VRHE (Ered)Expt.+DFTEC0.5 M KHCO3[27]
Nickel-TPP(-COOH)4NiN4CO, 82.8%−1.52 VFc+/Fc(Ered)/−1.29 VRHEExpt.+DFTPCCH3CN/H2O[35]
Nickel-21-oxa-TPP(-COOH)4NiN3OCO, 94.0%−1.41 VFc+/Fc(Ered)/−1.18 VRHE
Nickel-21-thia-TPP(-COOH)4NiN3SCO, 96.4%−0.82 VFc+/Fc(Ered)/−0.60 VRHE
Nickel-TPPNiN4CO, <5%/Expt.+DFTPCCH3CN[36]
Nickel-NCTPPNiN3CCO, 98%−1.41 VAg/Ag+(Ered)/−1.26 VRHE
Copper-TPPCuN4CO, <15%−1.8 VRHE(Ered)Expt.EC1 M KOH[89]
Copper-NCTPPCuN3CCH4, >60%−1.6 VRHE(Ered)
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

Li, Z.; Wei, Q.; Ren, Z.; Xie, J. Recent Progress in Heteroatom-Containing Metalloporphyrin-Based Catalysts for CO2 Reduction. Molecules 2025, 30, 2287. https://doi.org/10.3390/molecules30112287

AMA Style

Li Z, Wei Q, Ren Z, Xie J. Recent Progress in Heteroatom-Containing Metalloporphyrin-Based Catalysts for CO2 Reduction. Molecules. 2025; 30(11):2287. https://doi.org/10.3390/molecules30112287

Chicago/Turabian Style

Li, Zhuo, Qianqian Wei, Zhixin Ren, and Jing Xie. 2025. "Recent Progress in Heteroatom-Containing Metalloporphyrin-Based Catalysts for CO2 Reduction" Molecules 30, no. 11: 2287. https://doi.org/10.3390/molecules30112287

APA Style

Li, Z., Wei, Q., Ren, Z., & Xie, J. (2025). Recent Progress in Heteroatom-Containing Metalloporphyrin-Based Catalysts for CO2 Reduction. Molecules, 30(11), 2287. https://doi.org/10.3390/molecules30112287

Article Metrics

Back to TopTop