Next Article in Journal
Density Functional Theory Study in Photocatalytic Water Splitting via Covalent Triazine Frameworks Functioned by Benzothiophene Sulfone
Previous Article in Journal
Advanced Catalytic Processes for Wastewater Treatment
Previous Article in Special Issue
Sulfur Vacancy Engineering in Photocatalysts for CO2 Reduction: Mechanistic Insights and Material Design
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Copper Active Sites in Metal–Organic Frameworks Advance CO2 Adsorption and Photocatalytic Conversion

School of Chemistry and Chemical Engineering, Jiangsu University, Zhenjiang 212013, China
*
Authors to whom correspondence should be addressed.
Catalysts 2025, 15(9), 856; https://doi.org/10.3390/catal15090856
Submission received: 29 July 2025 / Revised: 26 August 2025 / Accepted: 2 September 2025 / Published: 4 September 2025
(This article belongs to the Special Issue Catalytic Carbon Emission Reduction and Conversion in the Environment)

Abstract

The photocatalytic reduction of CO2 into high-value chemicals utilizing solar energy represents a sustainable approach to mitigating greenhouse gas emissions and advancing renewable chemical production. Recently, copper-based metal–organic frameworks (Cu-MOFs) have been extensively researched for their potential in photocatalytic CO2 reduction, due to their high affinity for capturing CO2, the presence of unsaturated Cu sites, and their advantageous photochemical properties. In this review, we first provide an overview of Cu active sites in the secondary building units (SBUs) of MOFs, focusing on their selective adsorption of CO2 gas and analyzing the mechanisms of the multi-electron transfer processes involved in Cu-based photocatalytic reduction of CO2. Ultimately, this article outlines the existing obstacles and suggests potential avenues for future research.

Graphical Abstract

1. Introduction

The intensification of anthropogenic activities has led to a significant increase in atmospheric CO2 concentrations, exceeding 400 ppm, thereby accelerating global warming and climate change [1,2,3]. To mitigate these effects, research has concentrated on two primary strategies: physical sequestration of CO2 via geological storage methods and chemical conversion of CO2 into value-added chemicals and fuels through catalytic processes [4,5]. Chemical conversion encompasses direct electrochemical and photocatalytic reduction pathways, with the latter leveraging solar energy to drive the endothermic transformation of CO2 into simpler carbon-based molecules [6]. Photocatalytic reduction mimics natural photosynthesis, employing semiconductor photocatalysts to absorb photons and generate electron–hole pairs that facilitate the reduction of CO2 to C1 and C2 hydrocarbons and oxygenates, such as CO, CH4, CH3OH, C2H5OH, and HCHO [7,8]. Overcoming the kinetic and thermodynamic stability of CO2, characterized by a bond enthalpy of approximately +805 kJ mol−1 for the C=O bonds, remains a critical challenge [9]. Recent advances in photocatalysts focus on enhancing charge separation, expanding light absorption, and increasing the number of active sites to significantly improve the quantum efficiency of CO2 reduction under solar illumination [10,11,12]. This approach offers a sustainable pathway for renewable energy storage and greenhouse gas mitigation by converting atmospheric CO2 into useful chemical feedstocks and fuels with minimized energy input and environmental impact.
The photocatalytic reduction of CO2 involves sequential steps: photon absorption, charge carrier generation and separation, and CO2 adsorption and activation [13,14]. Upon photon absorption, electron–hole pairs are generated within the semiconductor photocatalyst. Effective separation of these charge carriers is vital to prevent recombination and improve the reduction efficiency [15]. Strategies such as forming heterojunctions or introducing electron traps enhance charge separation by creating internal electric fields or localized states that facilitate electron migration [16]. Concurrently, CO2 molecules adsorb onto the catalyst’s active sites, where increased surface area and exposed catalytic sites enhance adsorption energy and lower activation energy barriers, thereby accelerating reduction kinetics [17,18]. Despite progress in expanding the light absorption spectrum and reducing charge recombination [19], achieving high catalytic selectivity and overall efficiency remains challenging [20], necessitating the development of photocatalysts with broad spectral absorption, high charge mobility, abundant active sites, and robust CO2 adsorption capacity to advance sustainable solar-to-chemical energy conversion [21,22].
Metal–organic frameworks (MOFs) are crystalline materials composed of metal ions or clusters, named secondary building units (SBUs), coordinated with multi-dentate organic linkers, forming highly ordered, porous networks with exceptionally high specific surface areas, tunable pore dimensions, and abundant open metal sites [23,24,25]. These structural features confer advantages for a range of applications, including gas adsorption and separation, chemical sensing, catalysis, and CO2 capture and conversion [26,27,28]. Notably, MOFs exhibit remarkable potential as photocatalysts due to their broad-spectrum light absorption capabilities, which can be engineered through linker functionalization and metal node selection to optimize visible-light harvesting; their ability to facilitate the efficient separation and migration of photogenerated electron–hole pairs via electron-trapping centers; and the presence of uniformly distributed, accessible catalytically active sites, which enhance reaction kinetics and our mechanistic understanding of structure–activity relationships [29,30,31]. In addition to photocatalytic conversion, MOFs are employed in catalytic cycloaddition reactions and electrochemical reduction of CO2, leveraging their tunable pore environment, metal site density, and electronic properties to steer reaction selectivity and efficiency [32,33,34,35]. Consequently, the intrinsic structural versatility and functional tunability of MOFs make them highly promising platforms for advancing sustainable and efficient CO2 capture, activation, and conversion processes driven by solar energy.
Recently, copper-based metal–organic frameworks (Cu-MOFs) have shown remarkable efficacy in the photocatalytic reduction of CO2 [36,37,38,39]. The coordinatively unsaturated copper centers act as active catalytic sites, effectively adsorbing and activating CO2 molecules by lowering the activation energy required for reduction [40,41]. Their high surface area and accessible open copper sites further enhance CO2 adsorption, increasing the local concentration of reactants at active sites and thus boosting reduction rates [42]. Moreover, the organic linkers and metal centers work synergistically to expand the light absorption spectrum, enabling more efficient capture of solar energy and its conversion into chemical energy [43]. The combination of efficient light harvesting, effective charge separation, strong CO2 adsorption, and active copper sites positions Cu-MOFs as highly promising photocatalysts for CO2 reduction, paving the way for the conversion of solar energy into valuable chemical fuels.
To enable the strategic development of advanced Cu-MOFs for solar-driven CO2 reduction, it is important to deliver a thorough and up-to-date overview of the most recent scientific progress in this dynamic field. While several prior reviews have addressed aspects of MOF photocatalysis, there remains a need for an updated perspective that highlights new developments, offers novel insights, and discusses unresolved challenges [44,45,46,47]. In this review, we initially elucidate the fundamental mechanisms governing photocatalytic CO2 reduction, providing a comprehensive mechanistic framework. Building on this foundation, we highlight the critical importance of active site sensitivity and CO2 adsorption affinity in dictating catalytic performance. Subsequently, we systematically analyze the role of various copper (Cu) active sites in MOFs, including Cu (I), Cu (II), and single-atom Cu catalytic centers, in modulating CO2 adsorption characteristics. Based on these insights, we present a detailed mechanistic investigation of the multi-electron transfer pathways in Cu-MOF-catalyzed photocatalytic CO2 reduction. This includes an in-depth discussion of the two-electron, six-electron, eight-electron, and twelve-/fourteen-electron reduction processes, along with the identification and characterization of key reaction intermediates and their associated energetics. Finally, we critically assess the current limitations and unresolved challenges in MOF-based photocatalytic systems for CO2 reduction, and we discuss potential avenues for future research to enhance their performance and facilitate sustainable CO2 utilization.

2. Principles of MOF Photocatalysts for CO2 Conversion

Leveraging MOF-based photocatalysts to harness solar energy represents a promising approach for converting CO2 into valuable fuels and chemical commodities [48,49,50,51,52]. The process fundamentally involves the photoexcitation of electrons from the HOMO, primarily contributed by the organic linker, to the LUMO, mainly derived from the metal clusters, upon photon absorption [40]. This excitation generates separated charge carriers, with electrons in the LUMO and holes in the HOMO that must be effectively separated and transported to catalytic sites [53]. The photogenerated electrons in the LUMO diffuse to active metal centers within the MOFs, where they participate in multi-electron transfer reactions with adsorbed CO2 molecules, initiating reduction pathways (Figure 1). For efficient reduction, the LUMO must be thermodynamically aligned above the redox potentials of the desired reduction products; for example, the reduction of CO2 to CO requires a LUMO energy exceedingly approximately –0.104 V, while reductions to formic acid (~–0.171 V), methanol (~0.016 V), and methane (~0.169 V) demand correspondingly higher LUMO energies [54,55,56]. The tunability of MOFs, achieved by engineering organic linkers and metal nodes, allows precise modulation of their electronic structures, particularly their HOMO and LUMO, thereby optimizing light absorption, charge separation efficiency, and redox potentials [57]. For oxidation reactions occurring at the HOMO, the photogenerated holes typically oxidize water molecules or sacrificial agents (SAs), producing oxygen or other oxidation products.
Rational design strategies that emphasize ligand functionalization and the strategic selection of metal nodes can effectively enhance the photocatalytic performance of MOFs. Furthermore, MOFs inherently possess several intrinsic features that promote efficient CO2 reduction: (1) both organic linkers and unsaturated metal centers can serve as active catalytic sites; (2) their high surface area and porosity facilitate substantial CO2 adsorption, increasing the local CO2 concentration at active sites and lowering activation energy barriers; (3) the highly ordered porous architecture enables the incorporation of guest photoactive species, further augmenting photocatalytic activity [59,60,61,62,63,64]. Collectively, these properties underscore the immense potential of MOFs as versatile and highly tunable platforms for solar-driven CO2 conversion.

3. Cu Sites in MOFs for CO2 Adsorption

In photocatalytic CO2 reduction, the efficiency critically hinges on the catalyst’s ability to adsorb and activate CO2 molecules at active sites, which, in Cu-MOFs, is achieved through the presence of coordinatively unsaturated Cu centers acting as Lewis acid sites that form stable coordinate covalent bonds with the quadrupolar CO2, stabilizing it via electrostatic interactions between the electron-deficient metal centers and the partial negative charges on the oxygen atoms of CO2 [65,66,67,68,69,70]. The highly porous, high-surface-area framework provides an extensive internal volume, enabling physical adsorption and facilitating the diffusion of CO2 molecules; during synthesis, pore sizes can be precisely tuned to match the molecular dimensions of CO2 (~3.3 Å), promoting size-selective adsorption and minimizing the competitive binding of undesired gases [71]. Functional groups such as amino or hydroxyl groups incorporated into organic linkers introduce additional hydrogen-bonding and dipolar interactions with CO2, further increasing the binding energy and residence time—factors that are critical for activation and subsequent reduction [72]. The organic linkers extend light absorption into the visible spectrum, enhancing photon harvesting and charge generation, while electronic coupling between the metal centers and organic linkers promotes rapid charge transfer, minimizing electron–hole recombination and efficiently delivering electrons to adsorbed CO2 molecules [73]. The stable coordination environment maintains structural integrity under operational conditions, preventing framework degradation, while the electrostatic quadrupole moment of CO2 interacts favorably with the local electrostatic potential created by the open metal sites and functional groups, thus improving selectivity. Therefore, features such as accessible open metal sites, tailored pore structures, specific chemical functionalities, and extended light absorption synergistically enhance CO2 adsorption, activation, and reduction, underpinning the high catalytic performance of Cu-MOFs in solar-driven CO2 conversion.

3.1. Cu (I) Sites in MOFs

To address the challenges involved in developing an MOF-based integrated carbon capture and utilization material suitable for point sources in the cement industry, optimal MOFs should possess open metal sites (OMSs) for catalytic CO2 conversion, exhibit high stability under elevated temperatures and humid conditions, and selectively adsorb CO2 over other flue gas components. Cu(I)-MOFs are particularly promising due to their well-established catalytic capabilities for CO2 conversion, making them a strong candidate for this application. Sengupta et al. developed the Cu (I)-MOF NU-2100 (Cu2(BBTA), where H2BBTA = 1,5-Dihydrobenzo[1,2-d:4,5-d’]bis([1,2,3]triazole)) for integrated CO2 capture and conversion (Figure 2a). Under mild reaction conditions (50 °C and H2:CO2 = 3:1), NU-2100 demonstrated effective CO2 adsorption and catalytic conversion to formic acid with 100% selectivity (Figure 2b) [74]. Unlike in the as-synthesized MOF with acetonitrile molecules, the CO2 molecules in the activated form are preferentially localized at the vertices of the zigzag channels. As shown in Figure 2, this spatial distribution promotes weak C−H···OCO hydrogen-bonding interactions between the electron-deficient pyrazine and pyridine rings and the CO2 molecules, facilitating selective physisorption within the MOF’s confined environment [75]. This development represents a significant advancement toward the design of next-generation MOFs engineered for integrated carbon capture and conversion processes under industrial conditions.

3.2. Cu (II) Sites in MOFs

Although Cu (I)-based MOFs exhibit exceptional catalytic efficiency for CO2 conversion, their operational stability is frequently compromised by their susceptibility to oxidative degradation and hydrolysis under ambient air and moisture conditions. Hence, MOFs featuring coordinatively unsaturated Cu (II) centers demonstrate superior CO2 adsorption performance, driven by the availability of accessible open metal sites that enable strong, localized Lewis acid–base interactions and facilitate selective and reversible host–guest binding. A hexacarboxylate linker was developed and used to construct a Cu(II)-based MOF, known as MFM-160 (Figure 3a). The ligand structure, featuring a triazine backbone with a triangular configuration of isophthalic acid groups, was engineered to enhance CO2 uptake and enable the selective isolation of C2 hydrocarbons from methane (Figure 3b) [76]. These coordinating moieties facilitate the formation of an rht topology framework through integration with Cu(II) paddlewheel nodes. The rht topology has been extensively exploited in the development of MOFs with exceptional gas storage and separation performance, owing to its robust and accessible pore architecture. Furthermore, an amide-functionalized MOF, designated MFM-136 (Figure 3c), demonstrated a high CO2 adsorption capacity of 12.6 mmol g−1 at 20 bar and 298 K (Figure 3d). Notably, MFM-136 represents the first example of an acylamide pyrimidyl isophthalate MOF devoid of open metal sites, providing a unique system to elucidate guest–host interactions involving amide functionalities [77]. Neutron diffraction coupled with inelastic neutron scattering confirms that pendant amide functionalities do not establish direct coordination or hydrogen-bonding interactions with guest molecules such as CO2 or CH4. These findings highlight that functional group presence alone is insufficient for targeted guest affinity; instead, adsorption efficiency is predominantly governed by the spatial arrangement of the pore channels and the overall framework topology, which modulate host–guest interactions at the molecular level. Thus, gas adsorption performance results from complex interplay among the pore size, framework topology, and functional group chemistry. This underscores the importance of pore architecture and the chemical environment in collectively modulating guest affinity and selectivity in porous materials.

3.3. Cu Single-Atom Sites in MOFs

Single-atom Cu leverages maximal atomic utilization and distinctive electronic configurations to achieve enhanced catalytic activity and exceptional selectivity, thereby enabling the precise modulation of reaction pathways and facilitating the design of highly efficient, atomically precise catalytic systems. Mageed et al. engineered a single-atom copper catalyst supported on UiO-66, achieved by covalently anchoring individual copper atoms to defect sites in the zirconium-based framework [78]. Computational studies supported by experimental findings revealed that copper interacts with the MOF through hydroxyl and coordinated water ligands at defect sites on zirconium oxide clusters, contributing to the material’s stability and catalytic performance. The development of highly efficient single-atom-site electrocatalysts for the electrochemical reduction of carbon dioxide to methane has garnered significant interest for its potential use in intermittent renewable energy storage. However, realizing practical and selective catalytic systems remains a considerable challenge. Chen et al. developed N-heterocyclic carbene (NHC)-ligated copper single-atom sites (Cu SASs) embedded in a metal–organic framework (2Bn-Cu@UiO-67), shown in Figure 4c, which achieved an outstanding Faradaic efficiency (FE) of 81% for CO2 reduction to CH4 at @1.5 V vs. RHE with a current density of 420 mA cm−2. [79] The material’s porosity enhanced mass transport, promoting efficient diffusion of CO2 molecules to the active copper sites and thereby boosting catalytic accessibility. As Figure 4d shows, NU-1000 was functionalized with primary amino groups (–NH2), introduced via ligand modification, prior to MOF synthesis to enhance CO2 affinity, while post-synthetic solvent-assisted ligand exchange was employed to incorporate thiol (–SH) functionalities within the framework, providing affinity for copper species. This altered catalyst demonstrated complete selectivity for methanol synthesis at temperatures reaching 280 °C, with an approximate production rate of 100 milligrams of methanol per gram of catalyst per hour (Figure 4e) [80]. Additionally, the precisely characterized single-atom active sites of this catalyst will facilitate comprehensive mechanistic studies.

3.4. Pros and Cons of Cu (I, II, and Single-Atom)-Based MOFs

Cu(I)-based MOFs exhibit notable electronic properties, including high electrical conductivity, strong light absorption, and redox activity, making them promising candidates for photocatalysis, electrocatalysis, and electronic applications; however, their stability is often compromised under oxidative and humid conditions due to their susceptibility to oxidation to Cu(II), and their synthesis demands a strictly controlled inert atmosphere, which can limit scalability [74,75]. Cu(II)-based MOFs are distinguished by their exceptional thermal and chemical stability, diverse coordination geometries, and facile synthesis, enabling their widespread use in gas storage, separation, and heterogeneous catalysis; nevertheless, they generally possess lower electrical conductivity and limited redox flexibility, restricting their utility in applications requiring dynamic electron transfer [76,77]. Single-atom Cu MOFs leverage atomic dispersion to maximize catalytic site efficiency and selectivity, offering unique quantum confinement effects and electronic properties that enhance catalytic performance; however, the synthesis of stable, atomically dispersed Cu sites is technically challenging, often requiring sophisticated techniques to prevent the migration or aggregation of single atoms under operational conditions, which also complicates large-scale production and precise characterization [78,79,80].

4. Photocatalytic CO2 Reduction on Cu-MOFs

Previous studies have shown that copper-based metal–organic frameworks demonstrate exceptional sensitivity and affinity toward CO2 adsorption, owing to the presence of coordinatively unsaturated copper sites and tunable pore environments. However, the conversion of CO2 through this pathway requires several coordinated proton and electron transfer steps, leading to the formation of different intermediate species and making it challenging to steer the reaction toward a specific product [81]. In this section, we explore the photochemically initiated reactions that occur during both the two-electron and complex multi-electron reduction pathways of CO2.

4.1. Two-Electron Reduction Process

In photocatalytic CO2 reduction pathways involving two-electron transfer processes, the predominant reaction products are carbon monoxide (CO) and formic acid (HCOOH), resulting from the selective reduction of CO2 through controlled electron and proton transfer mechanisms. For CO2 reduction to CO, Cu-MOFs have been synthesized for photocatalytic reactions. As shown in Figure 5a, coordinated water molecules bound to the metal centers are readily displaced, exposing active metal sites where CO2 molecules can adsorb. Due to the lower LUMO energy level of MOF-Cu compared to [Ru(bpy)3]2+, photogenerated electrons in the LUMO of [Ru(bpy)3]2+ can be transferred to the surface of MOF-Cu. Subsequently, the adsorbed CO2 molecule accepts an electron at the active site, forming a radical CO2 intermediate. Through a proton-assisted two-electron transfer mechanism, this intermediate is ultimately reduced to carbon monoxide (CO). Finally, the excited state of the photosensitizer undergoes reductive quenching by the sacrificial electron donor TIPA, leading to the release of CO from the catalyst surface [82]. Feng et al. reported the design of an MOF called mPT-Cu/Re, which incorporates cuprous photosensitizers (Cu/Re-PSs) specifically engineered for CO2 reduction reactions (Figure 5b) [83]. Under 350−700 nm light irradiation, the mPT-Cu/Re catalyst facilitated continuous CO2 reduction to CO over a 48-hour period in a mixed solvent of N, N-dimethylacetamide (DMA) and water (v/v = 0.96:0.04), using BIH as the sacrificial electron donor. The gas chromatography analysis results, shown in Figure 5c, revealed a total turnover number (TON) of 1328 for CO production, with a CO to H2 selectivity ratio of 9:1.
A series of amine-functionalized Ti-based MOFs, including NH2-MIL-125(Ti), were synthesized using metalation, incorporation, and photodeposition techniques to introduce Cu and Ag into the framework. Among these, the post-synthetically metalated MOF with Cu2+ (MOF_met_0.5%Cu) demonstrated superior catalytic activity for CO2 reduction to HCOOH [84]. Additionally, single-site Co and Cu catalytically active centers were further encapsulated within the UiO-67 framework to facilitate directed and efficient electron transfer, thereby enhancing the overall efficiency of CO2 photoreduction (Figure 6a). The improved activity of UiO67-Ir-Cou 6/Co and UiO67-Ir-Cou 6/Cu is due to efficient electron flow from BIH to the excited MOF and fast intra-framework transfer to metal active sites, enhancing the CO2 reduction efficiency (Figure 6b) [85]. Cu incorporation enhances the photocatalytic performance of MOFs for CO2 reduction by increasing their CO2 adsorption capacity and decreasing their LUMO overpotentials, thereby facilitating more efficient charge transfer and activation of CO2 molecules.

4.2. Six-Electron Reduction Process

The six-electron conversion of CO2 to CH3OH demands a higher energy input and involves more complex challenges than the two-electron pathway. Goyal et al. demonstrated that the addition of metals to a material can modify its surface properties and enhance its overall efficiency [86]. Incorporated metal acts as a charge-carrier trap, boosting reaction rates and extending the lifespan of separated electron–hole pairs. During photocatalysis, electrons transition from lower to higher energy levels, forming a Schottky barrier at the metal–semiconductor interface. This barrier traps electrons, preventing their recombination with holes by blocking their migration back to the valence band. Visible light irradiation is predicted to excite electrons into the conduction band (CB) of Cu/ZIF-8 catalysts, facilitating charge separation. In contrast, pure ZIF-8 does not produce methanol, likely because CO2 reduction to methanol involves proton–electron transfer with a standard potential of E (CO2/CH3OH) = –0.38 V vs. NHE (Figure 7). Semiconductors are characterized by two key features: (1) the generation of a charge-separated state when the photon energy exceeds the bandgap, and (2) the mobility of charge carriers, which migrate away from their generation sites to minimize recombination and enable surface reactions.

4.3. Eight-Electron Reduction Process

The process of converting CO2 into CH4 and CH3COOH under solar illumination is an intricate multi-electron reaction pathway that includes numerous intermediate steps and yields different secondary products. Wu et al. engineered a surfactant-free method to encapsulate Cu2O nanowires within Cu3(BTC)2 metal–organic frameworks (BTC = 1,3,5-benzene tricarboxylate) [87]. The slow dissolution of Cu2+ from Cu2O, combined with the rapid nucleation of Cu3(BTC)2, resulted in the encapsulation of Cu2O nanowires within the Cu3(BTC)2 framework (Figure 8a). This process was visually indicated by color changes in CM, from reddish-brown to red and eventually to green, signaling the sequential growth of Cu2O and Cu3(BTC)2. Cu3(BTC)2 demonstrated an outstanding ability to catalytically reduce carbon dioxide into methane (Figure 8b). Jincai Zhao’s group developed a type II heterojunction photocatalyst by assembling two-dimensional materials through π–π interactions, combining graphitic carbon nitride with a copper-porphyrin-containing metal–organic framework. [88] The 2D-MOF component incorporated tetrakis(4-carboxyphenyl)porphyrin (TCPP) as the organic linker and was designated Cu-ZnTCPP. In this nomenclature, “Cu” refers to the copper node (Cu2(COO)4) in a paddle-wheel structure coordinated by carboxylate groups, while “Zn” denotes the central Zn ion (ZnN4) within the porphyrin ring (Figure 8c). As shown in Figure 8d, computational studies demonstrated that the Cu-ZnTCPP/g-C3N4 composite effectively stabilizes key intermediates in the stepwise reduction pathway, thereby promoting the selective conversion of CO2 to CO.

4.4. Twelve-/Fourteen-Electron Reduction Process

The photocatalytic reduction of CO2 to C2 hydrocarbons (C2H5OH and C2H4) is highly desirable due to its greater value. However, it remains a significant challenge owing to the complex multi-electron transfer process and the slow kinetics associated with C-C bond formation, which necessitate the coordinated action of multiple active sites working synergistically. Xu et al. co-immobilized Cu single-atom catalysts and Cu-doped Au nanoparticles on a mesoporous, photoactive MOF (ACN), creating a confined environment that combines multiple catalytic functionalities (Figure 9a) [89]. The detailed mechanism was elucidated through in situ DRIFTS analysis (Figure 9b). The increased absorption peaks at 1372 cm−1 and 1430 cm−1 were attributed to CO32− and HCO3 species, respectively, indicating CO2 adsorption on the surface of ACN. The characteristic peaks at 1646 cm−1 and 1737/1710 cm−1, corresponding to the key intermediates *COOH and *CHO, progressively intensified, providing direct evidence for the formation of CO and CH4. For C2 hydrocarbon formation, signals corresponding to intermediates *OCCO, *OCCHO, and *OCCHOH were observed at 1549/1531 cm−1, 1468 cm−1, and 1310 cm−1, respectively, confirming the occurrence of C–C coupling steps. In summary, the identification of these key C2 intermediates offers crucial mechanistic insights into the CO2 conversion pathway leading to C2 hydrocarbons.

4.5. Electronic and Geometric Factors of CO2 Conversion Mechanism

Product selectivity in photocatalytic CO2 reduction over Cu-MOFs is governed by a complex interplay of electronic and geometric factors at the active sites [90]. Electronically, the d-electron configuration of copper centers and their associated surface electronic states modulate the adsorption energies and stabilization of key reaction intermediates such as *COOH, *CO, and *CHO [91]. These intermediates’ binding affinities influence the reaction pathways, favoring the formation of specific products like CO, methane, or methanol. Geometrically, the local coordination environment—encompassing the surface facet orientation, atomic arrangement, defect sites, and pore architecture—dictates the adsorption modes and spatial orientation of reactants and intermediates [92]. Variations in active site geometry alter the transition state energies and facilitate or hinder particular reaction pathways, thereby steering selectivity toward different reduction products. Ultimately, the synergistic effects of the electronic structure and geometric configuration of Cu sites within the MOF matrix underpin mechanistic control over the product distribution in photocatalytic CO2 reduction.
The mechanistic pathways of photocatalytic CO2 reduction in Cu-MOFs display both commonalities and distinct differences when compared to other transition-metal-based MOFs, primarily arising from variations in their electronic structure, redox flexibility, and intermediate stabilization. In Cu-MOFs, the partially filled d-orbitals of copper facilitate efficient electron transfer and stabilize key reaction intermediates such as *COOH, *CO, and *CHO, thereby enabling multiple pathways that lead to a broad product spectrum including CO, methane, and methanol [93,94]. In contrast, transition metals like Co, Ni, and Fe possess different electronic configurations and redox potentials, which influence the adsorption energies and transition state barriers of intermediates, thereby biasing the reaction toward specific products—such as Fe-MOFs promoting multi-electron transfer pathways toward formate or other C1 products [95]. These differences in electronic properties and active site geometries modulate the stabilization of reaction intermediates and transition states, ultimately dictating the dominant mechanistic routes and product selectivity. A comprehensive understanding of these pathway distinctions provides critical insights for the rational design of MOFs with tailored active sites to optimize selectivity and efficiency in photocatalytic CO2 reduction.

5. Advanced Characterization

To achieve highly efficient photocatalysts, detailed characterization techniques are essential for unveiling critical processes such as light absorption, charge separation, and surface redox reactions. Numerous reviews have summarized the methodologies employed in photocatalysis research [96]. MOFs, with their atomically precise and tunable structures, offer a versatile platform for gaining deeper insights into photocatalytic mechanisms. In this context, advanced analytical methods become particularly valuable. This section first discusses single-crystal X-ray diffraction (SCXRD), which offers high-resolution, atomic-level insights into MOFs, enabling accurate determinations of connectivity, pore topology, and functionalization critical for elucidating structure–property relationships. Subsequently, X-ray absorption spectroscopy (XAS) provides element-specific, in situ insights into the electronic and local structural changes occurring in photocatalysts during operation, aiding in understanding their mechanisms and optimizing their performance. Finally, in situ diffuse reflectance infrared Fourier-transform spectroscopy (DRIFTS) facilitates the identification of reaction intermediates and provides mechanistic insights into catalytic processes.
Understanding and visualizing gas–adsorbent interactions is essential for developing improved materials for adsorption applications. As shown in Figure 10, structural analyses of N2, Ar, CO2, and C3H6 adsorption in Cu-MOF-1 under air-free conditions confirmed that gas exposure restored the larger pore phase, consistent with SPXRD results [97]. All gases except Ar interact with Cu(II) sites, revealing secondary (N2, CO2) and tertiary (N2) adsorption sites. Based on the SCXRD results, the Cu–gas distances indicate predominantly strong (N2, CO2, C3H6) and weaker (Ar, N2, CO2, C3H6) interactions. The gas loadings from SCXRD match the adsorption measurements, except that for Ar due to its higher disorder. Notably, the first single-crystal structure of a Cu–C3H6 adduct was obtained at 7.5 bar/290 K, with Cu–C distances of 2.33 Å, the shortest among the studied gases, confirming strong host–guest interactions. C3H6 preferentially forms single C3H6-Cu (II) complexes, interacting mainly with the framework rather than being centrally confined.

6. Summary

In summary, we initially delineated the fundamental mechanistic principles governing photocatalytic CO2 reduction, establishing a comprehensive theoretical framework that encompasses charge transfer dynamics, intermediate formation, and product selectivity. Building upon this foundation, we underscored the critical importance of active site sensitivity and CO2 adsorption affinity in modulating catalytic activity and selectivity. Subsequently, we systematically analyzed the role of various copper (Cu) active centers within MOFs—Cu (I), Cu (II), and single-atom Cu sites—influencing CO2 adsorption energies and activation barriers. Leveraging these insights, we presented an in-depth mechanistic investigation of the multi-electron transfer pathways involved in Cu-MOF-catalyzed photocatalytic CO2 reduction, encompassing two-electron, six-electron, eight-electron, and twelve-/fourteen-electron reduction processes. This analysis included a detailed characterization of key reaction intermediates, transition states, and their associated energetics, elucidating the thermodynamic and kinetic factors governing product formation. We discussed strategic directions for future research aimed at optimizing active site engineering, enhancing charge separation, and integrating these materials into scalable, sustainable CO2 conversion technologies.
Copper-based MOFs exhibit significant advantages in both their CO2 adsorption capacity (key performance indicators: uptake of CO2, adsorption isotherms) and photocatalytic reduction efficiency (key performance indicators: conversion rate, product selectivity) owing to their highly tunable pore structures and the presence of coordinatively unsaturated copper sites that promote strong CO2 affinity. The redox versatility of copper ions (Cu(I)/Cu(II)) facilitates efficient multi-electron transfer processes and the stabilization of reaction intermediates, thereby enhancing catalytic activity under visible light irradiation. Furthermore, the electronic structure of copper enables precise modulation of the MOF’s bandgap and light absorption properties, optimizing the charge separation and transfer essential for photocatalysis. As shown in Table 1, compared to other MOFs, such as Zr-, Co-, Ce-, Fe-, and Al-based frameworks, copper MOFs often demonstrate superior adsorption capacities and catalytic efficiencies, attributed to their adaptable coordination environments, high surface areas, and cost-effective synthesis. These combined attributes make copper-based MOFs particularly promising for scalable applications in CO2 capture and selective reduction, leveraging their structural and electronic tunability for enhanced performance.
Photocatalytic CO2 reduction using MOFs faces a multitude of interconnected challenges that hinder its potential for sustainable energy conversion. Many photocatalysts suffer from limited quantum efficiency and suboptimal light absorption spectra, restricting the generation and utilization of photogenerated charge carriers. Rapid charge recombination further diminishes the availability of electrons for reduction reactions, thereby impairing overall catalytic performance. Achieving high selectivity toward the desired products remains difficult due to competing reaction pathways and uncontrolled product distributions. The stability of MOF-based catalysts is often compromised by photocorrosion, structural degradation, and leaching under operational conditions, undermining their long-term durability. Additionally, the scarcity of catalytically active sites with sufficient affinity for CO2 adsorption and activation limits reaction rates, while the thermodynamic and kinetic barriers associated with multi-electron and multi-proton transfer steps result in high activation energies and sluggish kinetics. Mass transport limitations within porous structures hinder the efficient diffusion of CO2 to active sites and impede the removal of reduction products, further reducing efficiency. When considering practical deployment, the scalability of laboratory systems to industrial-scale applications requires thorough examination, alongside challenges related to reactor design, system integration, and maintaining catalyst stability under continuous solar irradiation. Long-term operational stability under real-world solar conditions, coupled with techno-economic factors such as material costs, energy input, and process optimization, is critical for commercial viability. Addressing these issues necessitates the development of advanced MOF materials with tailored electronic structures, enhanced light-harvesting capabilities, robust stability, and improved mass transport properties, complemented by innovative engineering solutions. Cu-MOFs, in particular, offer valuable mechanistic insights and serve as a foundational platform for the rational design of highly selective, durable, and economically feasible catalysts capable of advancing photocatalytic CO2 reduction toward practical, large-scale implementation.

Author Contributions

Conceptualization, E.J. and Y.Y. (Yan Yan); writing—original draft preparation, E.J.; writing—review and editing, Y.Y. (Yongsheng Yan) All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Postdoctoral Science Foundation of China (2023M741422) and Jiangsu Excellent Postdoc Project.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Jones, M.W.; Peters, G.P.; Gasser, T.; Andrew, R.M.; Schwingshackl, C.; Gütschow, J.; Houghton, R.A.; Friedlingstein, P.; Pongratz, J.; Quéré, C.L. National contributions to climate change due to historical emissions of carbon dioxide, methane, and nitrous oxide since 1850. Sci. Data 2023, 10, 1–23. [Google Scholar] [CrossRef]
  2. Fernández, J.; Sotenko, M.; Derevschikov, V.; Lysikov, A.; Rebrov, E.V. A radiofrequency heated reactor system for post-combustion carbon capture. Chem. Eng. Process. Process. Intensif. 2016, 108, 17–26. [Google Scholar] [CrossRef]
  3. Chronopoulos, T.; Fernandez-Diez, Y.; Maroto-Valer, M.M.; Ocone, R.; Reay, D.A. CO2 desorption via microwave heating for post-combustion carbon capture. Microporous Mesoporous Mater. 2014, 197, 288–290. [Google Scholar] [CrossRef]
  4. Nagireddi, S.; Agarwal, J.R.; Vedapuri, D. Carbon Dioxide Capture, Utilization, and Sequestration: Current Status, Challenges, and Future Prospects for Global Decarbonization. ACS Eng. Au 2023, 4, 22–48. [Google Scholar] [CrossRef]
  5. Geweda, A.E.; Zayed, M.E.; Khan, M.Y.; Alquaity, A.B.S. Mitigating CO2 emissions: A review on emerging technolo-gies/strategies for CO2 capture. J. Energy Inst. 2025, 118, 101911. [Google Scholar] [CrossRef]
  6. Kumagai, H.; Tamaki, Y.; Ishitani, O. Photocatalytic Systems for CO2 Reduction: Metal-Complex Photocatalysts and Their Hybrids with Photofunctional Solid Materials. Accounts Chem. Res. 2022, 55, 978–990. [Google Scholar] [CrossRef]
  7. Fang, S.; Rahaman, M.; Bharti, J.; Reisner, E.; Robert, M.; Ozin, G.A.; Hu, Y.H. Photocatalytic CO2 reduction. Nat. Rev. Methods Primers 2023, 3, 1–21. [Google Scholar] [CrossRef]
  8. Wu, H.; Han, D.; Hu, B.; Liang, J.; Tang, X.; Zhu, Z.; Huo, P. Insights into enhanced photocatalytic CO2 reduction on carbon nitride: A strategy of simultaneous O, S co-doping. J. Environ. Chem. Eng. 2025, 13, 116121. [Google Scholar] [CrossRef]
  9. Wang, J.; Jia, C.-S.; Li, C.-J.; Peng, X.-L.; Zhang, L.-H.; Liu, J.-Y. Thermodynamic Properties for Carbon Dioxide. ACS Omega 2019, 4, 19193–19198. [Google Scholar] [CrossRef]
  10. Yanagi, R.; Zhao, T.; Solanki, D.; Pan, Z.; Hu, S. Charge Separation in Photocatalysts: Mechanisms, Physical Parameters, and Design Principles. ACS Energy Lett. 2021, 7, 432–452. [Google Scholar] [CrossRef]
  11. Zhang, L.; Zhang, J.; Yu, J.; García, H. Charge-transfer dynamics in S-scheme photocatalyst. Nat. Reviews Chem. 2025, 9, 328–342. [Google Scholar] [CrossRef] [PubMed]
  12. Wei, T.; Zhou, J.; An, X. Recent advances in single-atom catalysts (SACs) for photocatalytic applications. Mater. Rep. Energy 2024, 4, 100285. [Google Scholar] [CrossRef]
  13. Lu, S.; Zhang, S.; Liu, Q.; Wang, W.; Hao, N.; Wang, Y.; Li, Z.; Luo, D. Recent advances in novel materials for photocatalytic carbon dioxide reduction. Carbon Neutralization 2024, 3, 142–168. [Google Scholar] [CrossRef]
  14. Kovačič, A.; Likozar, B.; Huš, M. Photocatalytic CO2 Reduction: A Review of Ab Initio Mechanism, Kinetics, and Multiscale Modeling Simulations. ACS Catal. 2020, 10, 14984–15007. [Google Scholar] [CrossRef]
  15. Kim, Y.; Smith, J.G.; Jain, P.K. Harvesting multiple electron–hole pairs generated through plasmonic excitation of Au nanoparticles. Nat. Chem. 2018, 10, 763–769. [Google Scholar] [CrossRef]
  16. Shen, W.; Qi, Q.; Hu, B.; Zhu, Z.; Huo, P.; Jiang, J.; Tang, X. Studying bimetals copper-indium for enhancing PCN photocatalytic CO2 reduction activity and selectivity mechanism. Ind. Eng. Chem. 2025, 145, 384–394. [Google Scholar] [CrossRef]
  17. Ye, Y.; Yang, H.; Qian, J.; Su, H.; Lee, K.-J.; Cheng, T.; Xiao, H.; Yano, J.; Goddard, W.A.; Crumlin, E.J. Dramatic differences in carbon dioxide adsorption and initial steps of reduction between silver and copper. Nat. Commun. 2019, 10, 1–9. [Google Scholar] [CrossRef]
  18. Álvarez, A.; Borges, M.; Corral-Pérez, J.J.; Olcina, J.G.; Hu, L.; Cornu, D.; Huang, R.; Stoian, D.; Urakawa, A. CO2 Activation over Catalytic Surfaces. ChemPhysChem 2017, 18, 3135–3141. [Google Scholar] [CrossRef]
  19. Han, C.; Kundu, B.K.; Liang, Y.; Sun, Y. Near-Infrared Light-Driven Photocatalysis with an Emphasis on Two-Photon Excitation: Concepts, Materials, and Applications. Adv. Mater. 2024, 36, 2307759. [Google Scholar] [CrossRef]
  20. Navarro-Jaén, S.; Virginie, M.; Bonin, J.; Robert, M.; Wojcieszak, R.; Khodakov, A.Y. Highlights and challenges in the selective reduction of carbon dioxide to methanol. Nat. Rev. Chem 2021, 5, 564–579. [Google Scholar] [CrossRef]
  21. Jiao, X.; Zheng, K.; Hu, Z.; Sun, Y.; Xie, Y. Broad-Spectral-Response Photocatalysts for CO2 Reduction. ACS Cent. Sci. 2020, 6, 653–660. [Google Scholar] [CrossRef] [PubMed]
  22. Zhu, S.; Wang, D. Photocatalysis: Basic Principles, Diverse Forms of Implementations and Emerging Scientific Opportunities. Adv. Energy Mater. 2017, 7, 1700841. [Google Scholar] [CrossRef]
  23. Furukawa, H.; Cordova, K.E.; O’kEeffe, M.; Yaghi, O.M. The Chemistry and Applications of Metal-Organic Frameworks. Science 2013, 341, 1230444. [Google Scholar] [CrossRef]
  24. Zhu, Z.; Xing, X.; Qi, Q.; Li, H.; Han, D.; Song, X.; Tang, X.; Ng, Y.H.; Huo, P. Regulation CN reduction of CO2 products selectivity by adjusting the number of V sites and mechanism exploration. Fuel 2025, 388, 134509. [Google Scholar] [CrossRef]
  25. Xu, W.; Pei, X.; Diercks, C.S.; Lyu, H.; Ji, Z.; Yaghi, O.M. A Metal-Organic Framework of Organic Vertices and Polyoxometalate Linkers as a Solid-State Electrolyte. J. Am. Chem. Soc. 2019, 141, 17522–17526. [Google Scholar] [CrossRef]
  26. Jiao, L.; Seow, J.Y.R.; Skinner, W.S.; Wang, Z.U.; Jiang, H.-L. Metal-organic frameworks: Structures and functional applications. Mater. Today 2019, 27, 43–68. [Google Scholar] [CrossRef]
  27. Senkovska, I.; Bon, V.; Mosberger, A.; Wang, Y.; Kaskel, S. Adsorption and Separation by Flexible MOFs. Adv. Mater. 2025, 2414724. [Google Scholar] [CrossRef]
  28. Li, D.; Yadav, A.; Zhou, H.; Roy, K.; Thanasekaran, P.; Lee, C. Advances and Applications of Metal-Organic Frameworks (MOFs) in Emerging Technologies: A Comprehensive Review. Glob. Challenges 2023, 8, 2300244. [Google Scholar] [CrossRef]
  29. Stanley, P.M.; Ramm, V.; Fischer, R.A.; Warnan, J. Analysis of metal-organic framework-based photosynthetic CO2 reduction. Nat. Synth. 2024, 3, 307–318. [Google Scholar] [CrossRef]
  30. Zhu, C.; Hou, J.; Wang, X.D.; Wang, S.; Xu, H.; Hu, J.; Jing, L.; Wang, S. Optimizing ligand-to-metal charge transfer in metal-organic frameworks to enhance photocatalytic performance. Chem. Eng. J. 2024, 499, 156527. [Google Scholar] [CrossRef]
  31. Behera, D.; Priyadarshini, P.; Parida, K. ZIF-8 metal–organic frameworks and their hybrid materials: Emerging photocatalysts for energy and environmental applications. Dalton Trans. 2025, 54, 2681–2708. [Google Scholar] [CrossRef] [PubMed]
  32. Stanley, P.M.; Haimerl, J.; Shustova, N.B.; Fischer, R.A.; Warnan, J. Merging molecular catalysts and metal-organic frameworks for photocatalytic fuel production. Nat. Chem. 2022, 14, 1342–1356. [Google Scholar] [CrossRef] [PubMed]
  33. Su, Q.; Guo, Q.; Wang, H.; Liu, M.; Zuo, C. Research progress of MOF-based materials in the photocatalytic CO2 reduction. Carbon Resour. Convers. 2024, 7, 100211. [Google Scholar] [CrossRef]
  34. Li, X.; Zhu, Q. MOF-based materials for photo-and electrocatalytic CO2 reduction. EnergyChem 2020, 2, 100033. [Google Scholar] [CrossRef]
  35. Li, T.; Wang, P.; He, M.; Zhang, T.; Yang, C.; Li, Z. Metal-organic frameworks for photocatalytical carbon dioxide reduction reaction. Coord. Chem. Rev. 2024, 521, 216179. [Google Scholar] [CrossRef]
  36. Zhou, Q.; Guo, Y.; Zhu, Y. Reticular copper dual sites embedded with semiconductor particles for selective CO2-to-C2H4 photoreduction. Nat. Catalysis 2025, 8, 728–739. [Google Scholar] [CrossRef]
  37. De Almeida, J.C.; Wang, Y.J.; Rodrigues, T.A.; Nunes, P.H.H.; De Mendonca, V.R.; Falsetti, P.H.E.; Savazi, L.V.; He, T.; Bardakova, A.V.; Rudakova, A.V.; et al. Copper-based Materials for Photo and Electrocatalytic Process: Advancing Renewable Energy and Environmental Applications. Adv. Funct. Mater. 2025, 2502901. [Google Scholar] [CrossRef]
  38. Liu, X.; Huang, S.; Li, G.; Chen, X.; Peng, J. Construction of molecular compartments on the HKUST-1 for space-limited enhancement of visible light CO2 reduction. J. Colloid Interface Sci. 2025, 690, 137347. [Google Scholar] [CrossRef]
  39. Liu, Y.Y.; Sun, Z.C.; Peng, C.; Wang, A.J. Microenvironment Regulation, Promoting CO2 Conversion to Mono-and Multi-carbon Products over Cu-Based Catalysts. Ind. Eng. Chem. Res. 2024, 63, 21613–21644. [Google Scholar] [CrossRef]
  40. Fan, L.; Li, F.; Liu, T.; Huang, J.E.; Miao, R.K.; Yu, Y.; Feng, S.; Tai, C.; Hung, S.; Tsai, H.-J.; et al. Atomic-level Cu active sites enable energy-efficient CO2 electroreduction to multicarbon products in strong acid. Nat. Synthesis 2024, 4, 262–270. [Google Scholar] [CrossRef]
  41. Li, Q.; Wu, J.; Yang, C.; Li, S.; Long, C.; Zhuang, Z.; Li, Q.; Guo, Z.; Huang, X.; Tang, Z.; et al. Ultralow Coordination Copper Sites Compartmentalized within Ordered Pores for Highly Efficient Electrosynthesis of n-Propanol from CO2. J. Am. Chem. Soc. 2025, 147, 6688–6697. [Google Scholar] [CrossRef] [PubMed]
  42. Cheng, D.; Zhao, Z.-J.; Zhang, G.; Yang, P.; Li, L.; Gao, H.; Liu, S.; Chang, X.; Chen, S.; Wang, T.; et al. The nature of active sites for carbon dioxide electroreduction over oxide-derived copper catalysts. Nat. Commun. 2021, 12, 1–8. [Google Scholar] [CrossRef] [PubMed]
  43. Ye, Q.; Cairnie, D.R.; Troya, D.; Kumar, N.; Yang, X.; Morris, A.J. Photoinduced Dynamic Ligation in Metal-Organic Frameworks. J. Am. Chem. Soc. 2023, 146, 101–105. [Google Scholar] [CrossRef]
  44. Li, D.; Kassymova, M.; Cai, X.; Zang, S.; Jiang, H. Photocatalytic CO2 reduction over metal-organic framework-based materials. Coord. Chem. Reviews 2020, 412, 213262. [Google Scholar] [CrossRef]
  45. Dhakshinamoorthy, A.; Li, Z.; Yang, S.; García, H. Metal-organic framework heterojunctions for photocatalysis. Chem. Soc. Reviews 2024, 53, 3002–3035. [Google Scholar] [CrossRef]
  46. Huang, N.-Y.; Zheng, Y.-T.; Chen, D.; Chen, Z.-Y.; Huang, C.-Z.; Xu, Q. Reticular framework materials for photocatalytic organic reactions. Chem. Soc. Rev. 2023, 52, 7949–8004. [Google Scholar] [CrossRef]
  47. Jin, H.; Zhao, P.; Qian, Y.; Xiao, J.; Chao, Z.; Jiang, H. Metal-organic frameworks for organic transformations by photocatalysis and photothermal catalysis. Chem. Soc. Reviews 2024, 53, 9378–9418. [Google Scholar] [CrossRef]
  48. Chang, X.-L.; Yan, T.; Pan, W.-G. Toward Tailoring Metal-Organic Frameworks for Photocatalytic Reduction of CO2 to Fuels. Cryst. Growth Des. 2024, 24, 2619–2644. [Google Scholar] [CrossRef]
  49. Huang, N.Y.; Li, B.; Wu, D.J.; Chen, Z.Y.; Shao, B.; Chen, D.; Zheng, Y.T.; Wang, W.J.; Yang, C.Z.; Gu, M.; et al. Crystal Engineering of MOF-Derived Bimetallic Oxide Solid Solution Anchored with Au Nanoparticles for Photocatalytic CO2 Reduction to Syngas and C2 Hydrocarbons. Angew. Chem. Int. Ed. 2024, 63, e202319177. [Google Scholar] [CrossRef]
  50. Huang, T.; Yang, H.; Xu, W.; Sun, Y.; Pang, H. Research progress of MOF-based materials in photocatalytic reduction of CO2 and N2. Chem. Catal. 2024, 4, 100211. [Google Scholar] [CrossRef]
  51. Song, B.; Song, W.T.; Liang, Y.H.; Liu, Y.; Li, B.W.; Li, H.; Zhang, L.; Ma, Y.H.; Ye, R.Q.; Tang, B.Z.; et al. Direct Synthesis of Topology-Controlled BODIPY and CO2-Based Zirconium Metal-Organic Frameworks for Efficient Photocatalytic CO2 Reduction. Angew. Chem. Int. Ed. 2025, 64, e202421248. [Google Scholar] [CrossRef]
  52. Wang, Y.J.; Zhang, X.; Sun, Y.T.; Chen, K.Z.; Zhao, H.Y.; Lin, Q.Q.; Li, R.; Li, J.R. Modular MOF-red phosphorus hetero-structure for synergistic capture and photocatalytic reduction of dilute CO2 into ethane. Appl. Catal. B Environ. 2025, 378, 125568. [Google Scholar] [CrossRef]
  53. Wu, Y.; Qu, Y.; Su, C.; Yang, X.; Yang, Y.; Zhang, Y.; Huang, W. Enhanced Photoinduced Carrier Separation in Fe-MOF-525/CdS for Photocatalytic Hydrogen Evolution: Improved Catalytic Dynamics with Specific Active Sites. Inorg. Chem. 2023, 62, 21290–21298. [Google Scholar] [CrossRef]
  54. Hamad, S.; Hernandez, N.C.; Aziz, A.; Ruiz-Salvador, A.R.; Calero, S.; Grau-Crespo, R. Electronic structure of porphyrin-based metal–organic frameworks and their suitability for solar fuel production photocatalysis. J. Mater. Chem. A 2015, 3, 23458–23465. [Google Scholar] [CrossRef]
  55. Liu, H.; Zhou, Z.; Xie, L.; Liu, C.; Cai, L.; Wu, X.; Liu, T. Delocalized Orbitals Over Metal Clusters and Organic Linkers Enable Boosted Charge Transfer in Metal-Organic Framework for Overall CO2 Photoreduction. Angew. Chemie 2024, 136, e202411508. [Google Scholar] [CrossRef]
  56. Li, K.; Peng, B.; Peng, T. Recent Advances in Heterogeneous Photocatalytic CO2 Conversion to Solar Fuels. ACS Catal. 2016, 6, 7485–7527. [Google Scholar] [CrossRef]
  57. Grau-Crespo, R.; Aziz, A.; Collins, A.W.; Crespo-Otero, R.; Hernández, N.C.; Rodríguez-Albelo, L.M.; Ruiz-Salvador, A.R.; Calero, S.; Hamad, S. Modelling a Linker Mix-and-Match Approach for Controlling the Optical Excitation Gaps and Band Alignment of Zeolitic Imidazolate Frameworks. Angew. Chemie 2016, 128, 16246–16250. [Google Scholar] [CrossRef]
  58. Sun, K.; Qian, Y.; Jiang, H.L. Metal-Organic Frameworks for Photocatalytic Water Splitting and CO2 Reduction. Angew. Chem. Int. Ed. Engl. 2023, 62, e202217565. [Google Scholar] [CrossRef]
  59. White, J.L.; Baruch, M.F.; Pander, J.E.; Hu, Y.; Fortmeyer, I.C.; Park, J.E.; Zhang, T.; Liao, K.; Gu, J.; Yan, Y.; et al. Light-Driven Heterogeneous Reduction of Carbon Dioxide: Photocatalysts and Photoelectrodes. Chem. Rev. 2015, 115, 12888–12935. [Google Scholar] [CrossRef]
  60. Linares, N.; Silvestre-Albero, A.M.; Serrano, E.; Silvestre-Albero, J.; García-Martínez, J. Mesoporous materials for clean energy technologies. Chem. Soc. Rev. 2014, 43, 7681–7717. [Google Scholar] [CrossRef]
  61. Karmakar, S.; Barman, S.; Rahimi, F.A.; Rambabu, D.; Nath, S.; Maji, T.K. Confining charge-transfer complex in a metal-organic framework for photocatalytic CO2 reduction in water. Nat. Commun. 2023, 14, 4508. [Google Scholar] [CrossRef] [PubMed]
  62. Liang, J.X.; Yu, H.; Shi, J.J.; Li, B.; Wu, L.X.; Wang, M. Dislocated Bilayer MOF Enables High-Selectivity Photocatalytic Reduction of CO2 to CO. Adv. Mater. 2023, 35, 2209814. [Google Scholar] [CrossRef]
  63. Liu, K.-G.; Bigdeli, F.; Panjehpour, A.; Larimi, A.; Morsali, A.; Dhakshinamoorthy, A.; Garcia, H. Metal organic framework composites for reduction of CO2. Coord. Chem. Rev. 2023, 493, 493. [Google Scholar] [CrossRef]
  64. Qian, Z.; Zhang, R.; Xiao, Y.; Huang, H.; Sun, Y.; Yang, C.; Ma, T.; Sun, X. Trace to the Source: Self-Tuning of MOF Photocatalysts. Adv. Energy Materials 2023, 13, 2300086. [Google Scholar] [CrossRef]
  65. Li, D.; Wang, K.; Li, J.; Li, Z.; Wang, H.; Wang, Y. Strategies for optimizing the efficiency and selectivity of photocatalytic aqueous CO2 reduction: Catalyst design and operating conditions. Nano Energ. 2024, 133, 110460. [Google Scholar] [CrossRef]
  66. Gao, C.-Y.; Tian, H.-R.; Ai, J.; Li, L.-J.; Dang, S.; Lan, Y.-Q.; Sun, Z.-M. A microporous Cu-MOF with optimized open metal sites and pore spaces for high gas storage and active chemical fixation of CO2. Chem. Commun. 2016, 52, 11147–11150. [Google Scholar] [CrossRef]
  67. Gupta, A.K.; Guha, N.; Krishnan, S.; Mathur, P.; Dhirendra, K. A Three-Dimensional Cu(II)-MOF with Lewis acid−base dual functional sites for Chemical Fixation of CO2 via Cyclic Carbonate Synthesis. J. CO2 Util. 2020, 39, 101173. [Google Scholar] [CrossRef]
  68. Stolar, T.; Prasnikar, A.; Martinez, V.; Karadeniz, B.; Bjelic, A.; Mali, G.; Friscic, T.; Likozar, B.; Uzarevic, K. Scalable Mech-anochemical Amorphization of Bimetallic Cu-Zn MOF-74 Catalyst for Selective CO2 Reduction Reaction to Methanol. ACS Appl. Mater. Interfaces 2021, 13, 3070–3077. [Google Scholar] [CrossRef] [PubMed]
  69. Tapiador, J.; Leo, P.; Gándara, F.; Calleja, G.; Orcajo, G. Robust Cu-URJC-8 with mixed ligands for mild CO2 cycloaddition reaction. J. CO2 Util. 2022, 64, 102166. [Google Scholar] [CrossRef]
  70. Zhou, G.; Chen, Y.; Chen, G.L.; Xu, H.; Yin, W.; Wang, B.; Zhu, X.; Ning, X.; Chu, P.K.; Wang, X.; et al. Enhanced CO2 Photoreduction in Pure Water Systems by Surface-Reconstructed Asymmetric Mn−Cu Sites. Appl. Catalysis B Environ. Energy 2024, 361, 124617. [Google Scholar] [CrossRef]
  71. Heine, T.; Dinca, M.; Zhu, G. Physical Phenomena in Porous Frameworks. Accounts Chem. Res. 2025, 58, 327–329. [Google Scholar] [CrossRef] [PubMed]
  72. Kim, D.; Kang, M.; Ha, H.; Hong, C.S.; Kim, M. Multiple functional groups in metal–organic frameworks and their positional regioisomerism. Coord. Chem. Rev. 2021, 438, 213892. [Google Scholar] [CrossRef]
  73. Ren, D.; Xia, H.; Zhou, K.; Wu, S.; Liu, X.; Wang, X.; Li, J. Tuning and Directing Energy Transfer in the Whole Visible Spectrum through Linker Installation in Metal–Organic Frameworks. Angew. Chem. Int. Ed. Engl. 2021, 60, 25048–25054. [Google Scholar] [CrossRef] [PubMed]
  74. Sengupta, D.; Bose, S.; Wang, X.; Schweitzer, N.M.; Malliakas, C.D.; Xie, H.; Duncan, J.; Kirlikovali, K.O.; Yildirim, T.; Farha, O.K. Integrated CO2 Capture and Conversion by a Robust Cu(I)-Based Metal-Organic Framework. J. Am. Chem. Soc. 2024, 146, 27006–27013. [Google Scholar] [CrossRef]
  75. Miller, R.G.; Warren, M.R.; Allan, D.R.; Brooker, S. Direct Crystallographic Observation of CO2 Captured in Zig Zag Channels of a Copper(I) Metal-Organic Framework. Inorg. Chem. 2020, 59, 6376–6381. [Google Scholar] [CrossRef]
  76. Trenholme, W.J.F.; Kolokolov, D.I.; Bound, M.; Argent, S.P.; Gould, J.A.; Li, J.; Barnett, S.A.; Blake, A.J.; Stepanov, A.G.; Besley, E.; et al. Selective Gas Uptake and Rotational Dynamics in a (3, 24)-Connected Metal–Organic Framework Material. J. Am. Chem. Soc. 2021, 143, 3348–3358. [Google Scholar] [CrossRef]
  77. Benson, O.; Da Silva, I.; Argent, S.P.; Cabot, R.; Savage, M.; Godfrey, H.G.W.; Yan, Y.; Parker, S.F.; Manuel, P.; Lennox, M.J.; et al. Amides Do Not Always Work: Observation of Guest Binding in an Amide–Functionalized Porous Metal-Organic Framework. J. Am. Chem. Soc. 2016, 138, 14828–14831. [Google Scholar] [CrossRef]
  78. Abdel-Mageed, A.M.; Rungtaweevoranit, B.; Parlińska-Wojtan, M.; Pei, X.; Yaghi, O.M.; Behm, R.J. Highly Active and Stable Single–Atom Cu Catalysts Supported by a Metal–Organic Framework. J. Am. Chem. Soc. 2019, 141, 5201–5210. [Google Scholar] [CrossRef]
  79. Chen, S.; Li, W.; Jiang, W.; Yang, J.; Zhu, J.; Wang, L.; Ou, H.; Zhuang, Z.; Chen, M.; Sun, X.; et al. MOF Encapsulating N-Heterocyclic Carbene-Ligated Copper Single-Atom Site Catalyst towards Efficient Methane Electrosynthesis. Angew. Chemie Int. Ed. 2021, 61, e202114450. [Google Scholar] [CrossRef]
  80. Rosado, A.; Popa, I.-M.; Markeb, A.A.; Moral-Vico, J.; Naughton, E.M.; Eckhardt, H.-G.; Ayllon, J.A.; López-Periago, A.M.; Domingo, C.; Negahdar, L. Multifunctionalized zirconium-based MOF as a novel support for dispersed copper: Application in CO2 adsorption and catalytic conversion. J. Mater. Chem. A 2024, 12, 21758–21771. [Google Scholar] [CrossRef]
  81. Zhang, Q.; Tsai, H.J.; Li, F.; Wei, Z.; He, Q.; Ding, J.; Liu, Y.; Lin, Z.-Y.; Yang, X.; Chen, Z.; et al. Boosting the Proton-coupled Electron Transfer via Fe−P Atomic Pair for Enhanced Electrochemical CO2 Reduction. Angew. Chem. Int. Ed. 2023, 62, e202311550. [Google Scholar] [CrossRef]
  82. Wang, X.-K.; Liu, J.; Zhang, L.; Dong, L.-Z.; Li, S.-L.; Kan, Y.-H.; Li, D.-S.; Lan, Y.-Q. Monometallic Catalytic Models Hosted in Stable Metal-Organic Frameworks for Tunable CO2 Photoreduction. ACS Catal. 2019, 9, 1726–1732. [Google Scholar] [CrossRef]
  83. Feng, X.; Pi, Y.; Song, Y.; Brzezinski, C.; Xu, Z.; Li, Z.; Lin, W. Metal-Organic Frameworks Significantly Enhance Photocatalytic Hydrogen Evolution and CO2 Reduction with Earth-Abundant Copper Photosensitizers. J Am Chem Soc 2020, 142, 690–695. [Google Scholar] [CrossRef] [PubMed]
  84. Baluk, M.A.; Pieczyńska, A.; Mazierski, P.; Kroczewska, M.; Nikiforow, K.; Mikołajczyk, A.; Dołżonek, J.; Łuczak, J.; Zaleska-Medynska, A. Effect of copper and silver modification of NH2-MIL-125(Ti) on the photoreduction of carbon dioxide to formic acid over this framework under visible-light irradiation. Appl. Catalysis B Environ. Energy 2024, 354, 124107. [Google Scholar] [CrossRef]
  85. Guo, S.; Kong, L.; Wang, P.; Yao, S.; Lu, T.; Zhang, Z. Switching Excited State Distribution of Metal-Organic Framework for Dramatically Boosting Photocatalysis. Angew. Chemie Int. Ed. 2022, 61, e202206193. [Google Scholar] [CrossRef]
  86. Goyal, S.; Shaharun, M.S.; Kait, C.F.; Abdullah, B.; Ameen, M. Photoreduction of Carbon Dioxide to Methanol over Copper Based Zeolitic Imidazolate Framework-8: A New Generation Photocatalyst. Catalysts 2018, 8, 1–19. [Google Scholar] [CrossRef]
  87. Wu, H.; Kong, X.Y.; Wen, X.; Chai, S.P.; Lovell, E.C.; Tang, J.; Ng, Y.H. Metal-Organic Framework Decorated Cuprous Oxide Nanowires for Long-lived Charges Applied in Selective Photocatalytic CO2 Reduction to CH4. Angew. Chem. Int. Ed. Engl. 2021, 60, 8455–8459. [Google Scholar] [CrossRef]
  88. Xie, S.; Deng, C.; Huang, Q.; Zhang, C.; Chen, C.; Zhao, J.; Sheng, H. Facilitated Photocatalytic CO2 Reduction in Aerobic Environment on a Copper-Porphyrin Metal-Organic Framework. Angew. Chem. Int. Ed. Engl. 2023, 62, e202216717. [Google Scholar] [CrossRef]
  89. Chen, D.; Chu, B.; Li, F.; Zheng, Y.T.; Lu, Y.; Shao, B.; Li, L.; Huang, N.Y.; Xu, Q. Synergistic Catalysis by Cu Single Atoms and Atomically Cu-Doped Au Nanoparticles in a Metal-Organic Framework for Photocatalytic CO2 Reduction to C2H6. J. Am. Chem. Soc. 2025, 147, 22705–22713. [Google Scholar] [CrossRef]
  90. Qin, C.; Li, X.; Wang, T.; Xu, Z.; Chen, K.-J.; Pan, F. Metal-Organic Frameworks-Based Copper Catalysts for CO2 Electroreduction Toward Multicarbon Products. Exploration 2025, 5, 270011. [Google Scholar] [CrossRef]
  91. Li, D.; Liu, J.; Wang, B.; Huang, C.; Chu, P.K. Progress in Cu-Based Catalyst Design for Sustained Electrocatalytic CO2 to C2+ Conversion. Adv. Sci. 2025, 12, 2416597. [Google Scholar] [CrossRef] [PubMed]
  92. Xiao, L.; Ji, S.; Chen, Y.; Wang, D. Synthetic strategies for MOF-based single-atom catalysts for photo-and electro-catalytic CO2 reduction. IScience 2022, 25, 104177. [Google Scholar] [CrossRef]
  93. Scatena, R.; Guntern, Y.T.; Macchi, P. Electron Density and Dielectric Properties of Highly Porous MOFs: Binding and Mobility of Guest Molecules in Cu3(BTC)2 and Zn3(BTC)2. J. Am. Chem. Soc. 2019, 141, 9382–9390. [Google Scholar] [CrossRef] [PubMed]
  94. Nwosu, U.; Siahrostami, S. Copper-based metal-organic frameworks for CO2 reduction: Selectivity trends, design paradigms, and perspectives. Catalysis Sci. Technol. 2023, 13, 3740–3761. [Google Scholar] [CrossRef]
  95. Wang, D.; Huang, R.; Liu, W.; Sun, D.; Li, Z. Fe-Based MOFs for Photocatalytic CO2 Reduction: Role of Coordination Un-saturated Sites and Dual Excitation Pathways. ACS Catal. 2014, 4, 4254–4260. [Google Scholar] [CrossRef]
  96. Zhang, L.; Ran, J.; Qiao, S.-Z.; Jaroniec, M. Characterization of semiconductor photocatalysts. Chem. Soc. Rev. 2019, 48, 5184–5206. [Google Scholar] [CrossRef]
  97. Mon, M.; Bruno, R.; Tiburcio, E.; Grau-Atienza, A.; Sepúlveda-Escribano, A.; Ramos-Fernandez, E.V.; Fuoco, A.; Esposito, E.; Monteleone, M.; Jansen, J.C.; et al. Efficient Gas Separation and Transport Mechanism in Rare Hemilabile Metal-Organic Framework. Chem. Mater. 2019, 31, 5856–5866. [Google Scholar] [CrossRef]
  98. Saadh, M.J.; Mustafa, M.A.; Altalbawy, F.M.A.; Kaur, M.; Sharma, R.; Juraev, N.; Ghazy, H.; Abdulwahid, A.S.; Muften, N.F.; Saud, H.R.; et al. Incorporation anthracene and Cu to NH2-Zr-UiO-67 metal-organic framework: Introducing the simul-taneous selectivity and efficiency in photocatalytic CO2 reduction to ethanol. J. Mol. Struct. 2024, 1318, 139329. [Google Scholar] [CrossRef]
  99. Batoo, K.M.; Ali, E.; Hussein, B.A.; Al-khalidi, A.; Altimari, U.S.; Hussain, S.; Kareem, S.H.; Abid, M.K.; Alawadi, A.; Ali, I. The cooperative performance of iodo and copper in a Zr-based UiO-67 metal-organic framework for highly selective photocatalytic CO2 reduction to methanol. J. Mol. Struct. 2024, 1307, 137927. [Google Scholar] [CrossRef]
  100. Liu, Y.; Lv, L.; Feng, H.; Liu, H.; Wang, Y. Improved CO2 Photoreduction Activity and Selectivity Using an Indirect Z-Scheme Heterojunction of ZIF-67 and Bi4O5Br2 Nanostructures. ACS Appl. Nano Mater. 2024, 7, 27846–27857. [Google Scholar] [CrossRef]
  101. Ling, J.; Zhou, A.; Wang, W.; Jia, X.; Ma, M.; Li, Y. One-Pot Method Synthesis of Bimetallic MgCu-MOF-74 and Its CO2 Adsorption under Visible Light. ACS Omega 2022, 7, 19920–19929. [Google Scholar] [CrossRef]
  102. Khalilzadeh, A.; Shariati, A. Fe-N-TiO2/CPO-Cu-27 nanocomposite for superior CO2 photoreduction performance under visible light irradiation. Sol. Energy 2019, 186, 166–174. [Google Scholar] [CrossRef]
  103. Li, J.; Li, K.; Du, J.; Yang, H.; Song, C.; Guo, X. Impact of transition metal incorporation on the photocatalytic CO2 reduction activity of polymeric carbon nitride. J. CO2 Util. 2022, 64, 101173. [Google Scholar] [CrossRef]
  104. Yong, C.; Lu, G.; Wang, X.; Shi, G.; Wang, Y.; Xie, X.; Sun, J. Alkali-Metal-Modified Al-PMOF for Enhanced CO2 Adsorption and Photocatalytic Reduction. ACS Appl. Eng. Mater. 2025, 3, 1513–1521. [Google Scholar] [CrossRef]
  105. Batoo, K.M.; Ali, E.A.; Hussein, S.A.; Chandra, S.; Abdulwahid, A.S.; Kareem, S.; Hussain, S.; Omran, A.A.; Abid, M.K.; Alawadi, A. UiO-67 metal-organic frameworks with dual amino/iodo functionalization, and mixed Zr/Ce clusters: Highly selective and efficient photocatalyst for CO2 transformation to methanol. J. Mol. Struct. 2024, 1312, 138492. [Google Scholar] [CrossRef]
  106. Ding, L.; Bai, F.; Borjigin, B.; Li, Y.; Li, H.; Wang, X. Embedding Cs2AgBiBr6 QDs into Ce-UiO-66-H to in situ construct a novel bifunctional material for capturing and photocatalytic reduction of CO2. Chem. Eng. J. 2022, 446, 137102. [Google Scholar] [CrossRef]
  107. Crake, A.; Christoforidis, K.C.; Kafizas, A.; Zafeiratos, S.; Petit, C. CO2 capture and photocatalytic reduction using bifunctional TiO2/MOF nanocomposites under UV–vis irradiation. Appl. Catal. B Environ. 2017, 210, 131–140. [Google Scholar] [CrossRef]
Figure 1. Visual overview of the main stages in photocatalytic CO2 reduction and their associated electrochemical potentials (V vs. RHE, pH = 0). [58] Copyright 2023 Wiley-VCH.
Figure 1. Visual overview of the main stages in photocatalytic CO2 reduction and their associated electrochemical potentials (V vs. RHE, pH = 0). [58] Copyright 2023 Wiley-VCH.
Catalysts 15 00856 g001
Figure 2. (a) The SCXRD structure of NU-2100, highlighting the pore channel and the active sites. (b) Gas selectivity of NU-2100 over ethane (red), ethylene (blue), argon (pink), nitrogen (green), carbon monoxide (CO, dark blue), and oxygen (purple) [74]. Copyright 2024 American Chemical Society. (c) Two perspectives illustrating the positioning of the CO2 molecule at site B within the confined cavity formed by adjacent py-pz ligands [75]. Copyright 2020 American Chemical Society.
Figure 2. (a) The SCXRD structure of NU-2100, highlighting the pore channel and the active sites. (b) Gas selectivity of NU-2100 over ethane (red), ethylene (blue), argon (pink), nitrogen (green), carbon monoxide (CO, dark blue), and oxygen (purple) [74]. Copyright 2024 American Chemical Society. (c) Two perspectives illustrating the positioning of the CO2 molecule at site B within the confined cavity formed by adjacent py-pz ligands [75]. Copyright 2020 American Chemical Society.
Catalysts 15 00856 g002
Figure 3. (a) A structural diagram illustrating the distinct cavities within MFM-160, with large colored spheres indicating the regions accessible to solvent molecules. Atom types are distinguished by color: cyan for copper, red for oxygen, blue for nitrogen, and gray for carbon. (b) The adsorption and desorption isotherms of CO2 and CH4 on MFM-160a are depicted over the pressure range of 0 to 20 bar, with filled triangles indicating adsorption and open triangles representing desorption processes [76]. Copyright 2021 American Chemical Society. (c) A schematic illustrating the ordered arrangement of macrocyclic cavities within MFM-136, featuring larger A-type cages (shown in blue) interspersed with smaller B-type cages (depicted in orange). (d) Measured and computationally modeled adsorption isotherms of CO2 on desolvated MFM-136 are presented for temperatures of 273 K and 298 [77]. Copyright 2016 American Chemical Society.
Figure 3. (a) A structural diagram illustrating the distinct cavities within MFM-160, with large colored spheres indicating the regions accessible to solvent molecules. Atom types are distinguished by color: cyan for copper, red for oxygen, blue for nitrogen, and gray for carbon. (b) The adsorption and desorption isotherms of CO2 and CH4 on MFM-160a are depicted over the pressure range of 0 to 20 bar, with filled triangles indicating adsorption and open triangles representing desorption processes [76]. Copyright 2021 American Chemical Society. (c) A schematic illustrating the ordered arrangement of macrocyclic cavities within MFM-136, featuring larger A-type cages (shown in blue) interspersed with smaller B-type cages (depicted in orange). (d) Measured and computationally modeled adsorption isotherms of CO2 on desolvated MFM-136 are presented for temperatures of 273 K and 298 [77]. Copyright 2016 American Chemical Society.
Catalysts 15 00856 g003
Figure 4. (a) The atomic arrangement of an ideal UiO-66 framework is depicted, with yellow spheres indicating the available voids within the structure. (b) Single-atom Cu in the UiO-66 framework [78]. Copyright 2019 American Chemical Society. (c) The modeling structure of 2Bn-Cu@UiO-67 [79]. Copyright 2021 Wiley-VCH. (d) A diagram illustrating the synthesis process of NU-1000-NH2/PrSH–Cu, highlighting the roles of the introduced functional groups. (e) The adsorption behavior of CO2 at 0 °C for the tested materials is displayed alongside that for the unmodified NU-1000 [80]. Copyright 2024 Royal Society of Chemistry.
Figure 4. (a) The atomic arrangement of an ideal UiO-66 framework is depicted, with yellow spheres indicating the available voids within the structure. (b) Single-atom Cu in the UiO-66 framework [78]. Copyright 2019 American Chemical Society. (c) The modeling structure of 2Bn-Cu@UiO-67 [79]. Copyright 2021 Wiley-VCH. (d) A diagram illustrating the synthesis process of NU-1000-NH2/PrSH–Cu, highlighting the roles of the introduced functional groups. (e) The adsorption behavior of CO2 at 0 °C for the tested materials is displayed alongside that for the unmodified NU-1000 [80]. Copyright 2024 Royal Society of Chemistry.
Catalysts 15 00856 g004
Figure 5. (a) An illustrative pathway depicting the photocatalytic process by which MOFs convert CO2 into CO. (b) The structures and bridging ligands of mPT-Cu/Co and mPT-Cu/Re MOFs are shown, where the proximity between the Cu photosensitizers and Co/Re catalysts enhances their electron transfer and catalytic efficiency [82]. Copyright 2019 American Chemical Society. (c) Time-dependent CO2RR TONs of mPT-Cu/Re versus homogeneous controls [83]. Copyright 2020 American Chemical Society.
Figure 5. (a) An illustrative pathway depicting the photocatalytic process by which MOFs convert CO2 into CO. (b) The structures and bridging ligands of mPT-Cu/Co and mPT-Cu/Re MOFs are shown, where the proximity between the Cu photosensitizers and Co/Re catalysts enhances their electron transfer and catalytic efficiency [82]. Copyright 2019 American Chemical Society. (c) Time-dependent CO2RR TONs of mPT-Cu/Re versus homogeneous controls [83]. Copyright 2020 American Chemical Society.
Catalysts 15 00856 g005
Figure 6. (a) Post-synthetic functionalization of UiO-67-Ir-ppy and UiO-67-Ir-Cou with Cu2+ and Co2+ ions. (b) Potential energy diagram of PHE and CO2 photoreduction (V vs. SCE) [85]. Copyright 2022 Wiley-VCH.
Figure 6. (a) Post-synthetic functionalization of UiO-67-Ir-ppy and UiO-67-Ir-Cou with Cu2+ and Co2+ ions. (b) Potential energy diagram of PHE and CO2 photoreduction (V vs. SCE) [85]. Copyright 2022 Wiley-VCH.
Catalysts 15 00856 g006
Figure 7. (a) A diagram depicting the process of CO2 reduction into CH3OH. (b) The mechanism of CO2 reduction into CH3OH based on Cu/ZIF-8 [86]. Copyright 2018 Molecular Diversity Preservation International.
Figure 7. (a) A diagram depicting the process of CO2 reduction into CH3OH. (b) The mechanism of CO2 reduction into CH3OH based on Cu/ZIF-8 [86]. Copyright 2018 Molecular Diversity Preservation International.
Catalysts 15 00856 g007
Figure 8. (a) An illustration of the synthesis of Cu2O@Cu3(BTC)2 core–shell nanowires via a surfactant-free topotactic conversion strategy for selective photocatalytic reduction of CO2 to methane. (b) The time course of photocatalytic CH4 production and the normalized activity (%) across cycles. No CO or alcohol products were detected in any samples [87]. Copyright 2020 Wiley-VCH. (c) The structure of the Cu-Zn-TCPP MOF. (d) A Gibbs free energy profile comparing the CO2 reduction and oxygen reduction reactions at LC-Cu and HO-Cu active sites [88]. Copyright 2023 Wiley-VCH.
Figure 8. (a) An illustration of the synthesis of Cu2O@Cu3(BTC)2 core–shell nanowires via a surfactant-free topotactic conversion strategy for selective photocatalytic reduction of CO2 to methane. (b) The time course of photocatalytic CH4 production and the normalized activity (%) across cycles. No CO or alcohol products were detected in any samples [87]. Copyright 2020 Wiley-VCH. (c) The structure of the Cu-Zn-TCPP MOF. (d) A Gibbs free energy profile comparing the CO2 reduction and oxygen reduction reactions at LC-Cu and HO-Cu active sites [88]. Copyright 2023 Wiley-VCH.
Catalysts 15 00856 g008
Figure 9. (a) A schematic illustration of the synthesis process of ACN, highlighting the corresponding CO2 conversion sites. (b) In situ FTIR spectra of photocatalytic CO2 reduction on CAN [89]. Copyright 2025 American Chemical Society.
Figure 9. (a) A schematic illustration of the synthesis process of ACN, highlighting the corresponding CO2 conversion sites. (b) In situ FTIR spectra of photocatalytic CO2 reduction on CAN [89]. Copyright 2025 American Chemical Society.
Catalysts 15 00856 g009
Figure 10. Single-crystal X-ray structures of host−guest complexes with (a) N2, (b) Ar, (c) CO2, and (d) C3H6. Copper(II) atoms are shown as spheres and sticks, ligands as sticks, and guest gases as spheres and surface-rendered sticks [97]. Copyright 2019 American Chemical Society.
Figure 10. Single-crystal X-ray structures of host−guest complexes with (a) N2, (b) Ar, (c) CO2, and (d) C3H6. Copper(II) atoms are shown as spheres and sticks, ligands as sticks, and guest gases as spheres and surface-rendered sticks [97]. Copyright 2019 American Chemical Society.
Catalysts 15 00856 g010
Table 1. Adsorption and photocatalytic CO2 reduction capacity of MOF-based systems. TEOA: triethanolamine; BIH: 1,3 dimethyl-2-phenyl-2,3-dihydro-1H-benzo[d]imidazole.
Table 1. Adsorption and photocatalytic CO2 reduction capacity of MOF-based systems. TEOA: triethanolamine; BIH: 1,3 dimethyl-2-phenyl-2,3-dihydro-1H-benzo[d]imidazole.
MaterialsCO2 CapacityConditions for AdsorptionPhotocatalytic Conversion Product (μmol/g/h)Conditions for PhotocatalysisRef.
Cu/An
(1:2)@NZU67
71.6 cm3 g−1298 K
1 bar
EtOH: 624.17
Selectivity: 100%
Light: 450 nm,
TEOA (0.3 M), CH3CN/H2O
(4:1 v/v)
[98]
Cu/I2-Zr-BPDC/BPyDC48.6 cm3 g−1298 K
1 bar
MeOH: 9.4Light: 450 nm,
Ru(Bipy)3Cl2 (4 mg), TEOA (0.3 M)
[99]
Cu-Zr-BPyDC31.1 cm3 g−1MeOH: 1.88
Zr-I2-BPDC28.7 cm3 g−1/
Zr-BPyDC27.3 cm3 g−1/
1% ZIF-67CoCu(1:1)/
Bi4O5Br2
4.22 cm3 g−1273 K
750 mmHg
CO: 6469.88
Selectivity: 97%
UV−visible light, TEOA (0.3 M), BIH (10 mg), CH3CN (4 mL)
H2O (1 mL)
[100]
Mg0.4Cu0.6-MOF-744.58 mmol·g−1298 K
1 bar
CO: 6.18Visible light,
H2O (2 mL)
[101]
Fe-N-TiO2/CPO-Cu1.07 mmol·g−1298 K
1 bar
CH4: 47.52
MeOH: 2.17
Light: 350–600 nm,
H2O
[102]
CuCN7.4 cm3 g−1298 K
1 bar
CO: 246
Selectivity: 88%
Visible light,
TEOA (1 mL), bipyridine (100 μL), CH3CN (4 mL), CoCl2 (2 μmol), H2O (1 mL)
[103]
1-K-Al-PMOF3.46 mmol·g−1273.15 K
1 bar
CO: 587.2Visible light,
CH3CN/H2O/TEOA
(8:2:1 v/v)
[104]
Zr/Ce29%-UiO-67(NH 2, I)78.4 cm3 g−1298 K
1 bar
MeOH: 44.7
Selectivity: 100%
Light: 450 nm,
TEOA (0.3 M), CH3CN/H2O
(4:1 v/v)
[105]
20Cs2AgBiBr6/Ce-UiO-66-H58.8 cm3 g−1298 K
1 bar
CO: 309.01300 W Xe lamp,
H2O (50 mL)
[106]
NH2-UiO-661.3 mmol·g−1298 K
1 bar
CO: 1.5UV−visible light, gas/solid reactor[107]
1-TiMOF0.32 mmol·g−1298 K
1 bar
CO: 3.74
2-TiMOF0.36 mmol·g−1298 K
1 bar
CO: 4.24
3-TiMOF0.47 mmol·g−1298 K
1 bar
CO: 3.37
4-TiMOF0.56 mmol·g−1298 K
1 bar
CO: 2.85
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

Jiang, E.; Yan, Y.; Yan, Y. Copper Active Sites in Metal–Organic Frameworks Advance CO2 Adsorption and Photocatalytic Conversion. Catalysts 2025, 15, 856. https://doi.org/10.3390/catal15090856

AMA Style

Jiang E, Yan Y, Yan Y. Copper Active Sites in Metal–Organic Frameworks Advance CO2 Adsorption and Photocatalytic Conversion. Catalysts. 2025; 15(9):856. https://doi.org/10.3390/catal15090856

Chicago/Turabian Style

Jiang, Enhui, Yan Yan, and Yongsheng Yan. 2025. "Copper Active Sites in Metal–Organic Frameworks Advance CO2 Adsorption and Photocatalytic Conversion" Catalysts 15, no. 9: 856. https://doi.org/10.3390/catal15090856

APA Style

Jiang, E., Yan, Y., & Yan, Y. (2025). Copper Active Sites in Metal–Organic Frameworks Advance CO2 Adsorption and Photocatalytic Conversion. Catalysts, 15(9), 856. https://doi.org/10.3390/catal15090856

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop