金属有机骨架(MOFs)在光催化CO2环加成反应中的应用综述
English
Application of metal-organic frameworks (MOFs) in photocatalytic CO2 cycloaddition reaction: A mini review
-
0. Introduction
Over-reliance on fossil fuels has driven a persistent increase in global CO2 levels, amplifying the greenhouse effect and deepening ecological challenges[1-2]. Consequently, the utilization of CO2 as a resource has emerged as a pivotal challenge in achieving carbon neutrality goals. CO2, serving as a C1 resource, can be catalytically transformed into value-added chemicals such as methane, methanol, and cyclic carbonates[3-4]. Although the conversion of CO2 into cyclic carbonates achieves 100% atom utilization[5], traditional synthesis methods require high-temperature and high-pressure conditions, presenting significant challenges including excessive energy consumption and limited sustainability[6].
To overcome these constraints, photocatalytic technology utilizes solar energy to facilitate the moderate cycloaddition of CO2 with epoxides[7]. Upon photoexcitation with energy exceeding the bandgap (Eg), photoinduced electron-hole pairs are generated in the semiconductor, which would migrate to the surface active sites and undergo further reactions[8].
This manuscript focuses on the analysis of MOF-based photocatalysts for the CO2 cycloaddition reaction, with a particular emphasis on mechanistic insights and strategies for performance enhancement. The study elucidates the fundamental pathways governing radical-anion formation and epoxide activation, substantiated by empirical evidence. Furthermore, we examine three pivotal approaches, including ligand engineering, composite construction, and dopant incorporation, to address current limitations. This research endeavors to provide a critical overview of MOFs in the photocatalytic synthesis of cyclic carbonates, encompassing their structural characteristics, photocatalytic mechanisms, current challenges, and prospective developments.
1. Mechanism of photocatalytic CO2 cycloaddition
The excitation and separation of photogenerated carriers and their cooperative activation process with the reactants form the basis of the fundamental mechanism of the photocatalytic CO2 cycloaddition reaction. Current research focuses on the following two types of pathways:
(1) Radical-anion pathway: the electrons (e-) generated by photoexcitation migrate to the surface active sites through the directional transfer channel from MOFs′ ligands to metal clusters. The photoelectrons reduce CO2 to obtain ·CO2- radical anion, which then interacts with epoxide activated by the Lewis acid sites to produce cyclic carbonate[9] (Fig. 1a). The mechanistic pathway is exemplified through Fe-FcDC, wherein the organic ligand (1, 1′-ferrocenedicarboxylic acid, H2FcDC) facilitates photon absorption and charge carrier generation. The photogenerated electrons undergo directional migration through the synergistic interplay between FeO6 clusters and ferrocenyl moieties, ultimately reaching surface active sites where they facilitate CO2 activation to form ·CO2-, thus promoting the desired transformation[10].
Figure 1
Figure 1. Two mechanism diagrams of the photocatalytic CO2 cycloaddition reaction: (a) radical-anion pathway and (b) epoxide-activation pathway(2) Epoxide-activation pathway: alternatively, a negatively charged intermediate is created when the photoelectrons directly interact with the epoxide. The co-catalyst, most commonly tetrabutylammonium bromide (TBAB), facilitates Br--mediated nucleophilic ring opening, subsequent CO2 insertion, and ring closure to afford cyclic carbonate[11] (Fig. 1b). Other nucleophilic salts, such as tetrabutylammonium iodide (TBAI) or potassium iodide (KI), can also be employed. However, their efficacy may vary depending on the catalytic system[5]. A representative illustration is demonstrated by PB-modified WO3-x, wherein photoexcitation induces electron capture at oxygen vacancy sites. These trapped electrons subsequently interact with epoxy functionalities, generating highly reactive anionic intermediates that undergo successive transformations to yield the desired products[12].
2. Enhancement of MOFs′ photocatalytic properties
MOFs have emerged as promising photocatalysts for CO2 conversion into cyclic carbonates, owing to their unique structural advantages over conventional catalysts[13]. Their tunable pore architectures and abundant Lewis acid-base sites enable synergistic activation of both CO2 and epoxides, effectively lowering the energy barrier for ring-opening reactions. The photocatalytic mechanism in MOFs originates from light absorption by conjugated organic ligands, which generates excitons that undergo directional charge transfer to metal clusters through well-defined coordination pathways[14-15]. However, the practical application of MOF-based photocatalysts faces two fundamental challenges: narrow photo-response range and inefficient charge carrier separation[16]. This review aims to provide a comprehensive overview of recent progress in MOF-based photocatalytic systems for CO2 cycloaddition, systematically elucidating the structure-activity relationship in MOF-based photocatalysts, with particular focus on innovative design strategies to address these limitations, including extending the photo-response range through ligand engineering and sensitization, and enhancing charge separation efficiency via structural modifications and heterojunction construction. By comprehensively analyzing these approaches, we establish guidelines for developing high-performance MOF-based photocatalysts for sustainable CO2 utilization.
Analysis of the performance data in Table 1 reveals several key trends. Significant variation in yield is observed, ranging from below 0.2 mmol·g-1·h-1 to nearly 298 mmol·g-1·h-1, highlighting the profound impact of catalyst design. Factors influencing performance include the nature of the MOF (metal nodes, organic ligands, presence of dopants), the formation of composites or heterojunctions, light source intensity and wavelength, reaction time, and the specific epoxide substrate used (e.g., styrene oxide often requires longer times than epichlorohydrin). Notably, strategies such as incorporating redox-active ligands (e.g., Fe-FcDC[10]), heteroatom doping (e.g., UiO-67-B[42]), and constructing advanced composites (e.g., PMo12@Zr-Fc[41]) have successfully yielded catalysts exhibiting high productivity (> 40 mmol·g-1·h-1), demonstrating the effectiveness of the optimization approaches discussed in this review.
Table 1
Catalyst Light source Time /h Reaction substrate Conversion /% Yield /(mmol·g-1·h-1) Regulatory strategy Ref. Fe-FcDC 300 W xenon lamp 4 Epichlorohydrin 7 44 Ligand engineering [10] Co-PMOF 3 300 W xenon lamp 6 Epichlorohydrin > 99 61 Ligand engineering [17] MOF-997 300 W xenon lamp 24 Styrene oxide > 99 2 Ligand engineering [18] Th-IHEP-5 45 W fluorescent light 48 Styrene oxide 71 0.6 Ligand engineering [19] Zr-TTF-L1 300 W xenon lamp 8 Propylene oxide 45 14 Ligand engineering [20] MOF 1 (Mn) LED 18×3W 24 Epichlorohydrin 100 17 Ligand engineering [21] Fe-DBP(Co) 1000 W xenon lamp 12 Epichlorohydrin 97 103 Ligand engineering [22] FeTPyP 300 W xenon lamp 5 Styrene oxide 30 106 Ligand engineering [23] Bi-HHTP 300 W xenon lamp 4 Styrene oxide 34 112 Ligand engineering [24] Mg-MOF-74 400 mW·cm-2 12 1-Bromo-2,3-epoxypropane 97 5 Ligand engineering [25] UAEU-50 400 W halogen lamp 24 Styrene oxide 90 6 Ligand engineering [26] Co-HOF 400 W halogen lamp 24 Styrene oxide > 99 6 Ligand engineering [27] NH2-MIL-68 300 W xenon lamp 12 Styrene oxide 94 56 Ligand engineering [28] Ti-H2TPDC 8 W LED light 4 Epichlorohydrin 94 59 Ligand engineering [29] PB-modified WO3-x 300 W xenon lamp 8 Styrene oxide 94 171 Composite construction [12] W18O49/NH2-UiO-66 300 W xenon lamp 4 Styrene oxide 82 58 Composite construction [30] FeNbO4/NH2-MIL-125(Ti) 500 W halogen lamp 72 Propylene oxide 52 0.2 Composite construction [31] MIL-125@ZIF-67 3 W blue LED 36 1-Bromo-2,3-epoxypropane 99 0.3 Composite construction [32] Co2N0.67@ZIF-67 350 mW·cm-2 3 Epichlorohydrin 95 3 Composite construction [33] Zn20Fe1-ZIF/MXene 350 mW·cm-2 6 Epichlorohydrin 96 16 Composite construction [34] CMS@MIL-88-NH2 300 W xenon lamp 6 1, 2-Epoxybutane 72 24 Composite construction [35] Zn SA-NC 300 mW·cm-2 16 Epichlorohydrin 99 0.5 Composite construction [36] CoNHPC/TM-8 350 mW·cm-2 12 Styrene oxide 99 8 Composite construction [37] CuS@HKUST-1 Blue light 18 2-Chloromethyl-oxriane 98 11 Composite construction [38] Ag0.1/NWU-Cd3a 300 W xenon lamp 6 Styrene oxide 94 22 Composite construction [39] Zn-N-HOPCPs 350 mW·cm-2 10 Propylene oxide 95 29 Composite construction [40] PMo12@Zr-Fc 500 mW·cm-2 8 Epichlorohydrin 96 298 Composite construction [41] UiO-67-B 300 W xenon lamp 6 Propylene epoxide 6 45 Dopant incorporation [42] Bi-PCN-224 300 W xenon lamp 6 Propylene epoxide 99 5 Dopant incorporation [43] Tb-DBP(Co) 300 W xenon lamp 10 Epichlorohydrin > 99 127 Dopant incorporation [44] SNNU-97-InV 300 W xenon lamp 24 2-Bromooxirane 97 14 Dopant incorporation [45] Zn-UiO-bpydc 300 W xenon lamp 9 Propylene oxide 75 0.3 Dopant incorporation [46] IHEP-9 45 W compact fluorescent lamp 12 Styrene oxide 99 2 Dopant incorporation [47] 2.1 Ligand engineering
The photocatalytic efficiency of MOFs is intrinsically linked to the photoresponsive behavior of their organic ligands, which governs light absorption, charge generation, and carrier transport[48]. Recent advances in MOF design have focused on π-conjugation extension, post-synthetic ligand modification, and charge-transfer pathway engineering to optimize their photophysical properties.
To maximize solar energy utilization, extended π-conjugated architectures have been deliberately constructed in MOF ligands. A notable example is porphyrin-based Co-PMOFs, which integrate porphyrin and anthracene moieties to achieve panchromatic absorption (200-800 nm)[17]. This design enables near-complete epoxide conversion (> 99%) within 6 h under ambient conditions while maintaining exceptional cyclic stability. The enhanced light-harvesting capability stems from the synergistic effect of dual π-conjugated systems, which promote efficient exciton generation and charge separation.
Beyond the design of organic ligands themselves, precise modulation of the metal centers within macrocyclic ligand structures, such as porphyrins, represents a powerful sub-strategy under the umbrella of ligand engineering. Inspired by the molecular structure of chlorophyll, Das et al. integrated Mg2+ ions into the porphyrin ring to construct the PCN-224(Zr) catalyst. Under visible light irradiation, this material achieves 99% epoxide conversion, a 2.1-fold enhancement over the unmodified PCN-224 (47%) (Fig. 2a)[49]. This case highlights how engineering the metal node of a ligand unit itself can dramatically tune the electronic structure and catalytic activity, showcasing an advanced facet of ligand engineering[50].
Figure 2
Atomic-level control over MOFs′ catalytic sites can be achieved through post-synthetic ligand modification. A titanium-based MOF-901(Ti) was selectively oxidized at its imine bonds (C=N) to form amide linkages (—HN—CO—), yielding the novel MOF-997(Ti) (Fig. 2b)[18]. The introduced amide groups establish donor-acceptor active sites with 1.8 eV higher CO2 binding energy, resulting in a 99% cyclic carbonate yield, which is a 3.3-fold improvement over the parent MOF-901. While the fundamental light absorption onset might not shift dramatically, this electronic modification optimizes the utilization of photogenerated charges for the catalytic cycle, leading to a dramatic increase in yield. This approach demonstrates how subtle ligand alterations can fine-tune MOF electronic structures for superior photocatalytic performance.
To mitigate charge recombination, redox-active ligands have been employed to construct multistage carrier transport channels. Zhang et al. developed Fe-FcDC, a ferrocene-based MOF featuring H2FcDC ligands coordinated with FeO6 clusters (Fig. 2c)[10]. Unlike conventional organic linkers, FcDC introduces a metal-ligand-metal charge transfer pathway, which extends carrier migration distance and enhances photocurrent responsiveness. This design significantly improves photocatalytic CO2 cycloaddition efficiency by facilitating long-range electron transfer and suppressing recombination losses.
Thus, ligand engineering, through precise molecular design, stands as one of the most direct strategies to enhance the photocatalytic CO2 cycloaddition performance of MOFs. By expanding the π-conjugated system, conducting post-synthetic modifications, and introducing redox-active units, it is possible to effectively broaden the light response range, promote charge separation and transfer, and optimize the adsorption and activation processes of reactants on active sites. However, the synthesis of complex ligands often involves cumbersome steps, low yields, and high costs, and highly conjugated or functionalized ligands may encounter photochemical stability issues under prolonged illumination.
2.2 Composite construction
Composite construction is the most effective and widely used way to solve the problems of insufficient active sites, poor charge separation, and limited mass transfer of MOFs to enhance the catalytic efficiency. Currently, three main strategies are employed: (1) interface engineering to optimize charge separation, (2) pore structure modulation to improve mass transfer efficiency, and (3) multicomponent synergy to enhance catalytic performance.
In terms of interface design, molecular-level dispersion of Keggin-type phospho-molybdate (PMo12) within the pores of zirconium ferrocene MOF (Zr-Fc) nanosheets forms the PMo12@Zr-Fc composite (Fig. 3a)[41]. This design enhances the electronic interaction between PMo12 and the framework by utilizing the confinement effect of MOF pores, while retaining the high specific surface area and photoresponsiveness of Zr-Fc. It significantly increases the density of acid sites and optimizes cyclic stability through electrostatic binding, exhibiting a maximum cyclic carbonate yield of 298 mmol·g-1·h-1. Another typical example is the construction of W18O49/NH2-UiO-66 type Ⅱ heterojunction using the in situ solvothermal method, where the interface charge directed transport and multi-active site synergy in photocatalytic reactions (Fig. 3b)[30]. In the photocatalytic CO2 cycloaddition reaction, the composite catalyst exhibited 82% conversion of styrene oxide with 58 mmol·g-1·h-1 yield of cyclic carbonate, which were 3.4 and 7.5 times greater than those of W18O49 and NH2-UiO-66, respectively. NH2-MIL-125(Ti) and FeNbO4 constructed a heterojunction system with synergistic optimization of photon capture and carrier management through chemical modification and interface engineering, which achieves doubled cyclic carbonate yield compared to a single component[31]. MIL-125@ZIF-67 S-scheme heterojunction catalysts, where the built-in electric field formed at the interface effectively suppresses the recombination of photogenerated carriers, achieving a final catalytic efficiency of 99% (Fig. 3c)[32].
Figure 3
Figure 3. Mechanism diagrams of photocatalytic cyclic carbonates production: (a) PMo12@Zr-Fc (Copyrighted from Ref. [41] with a license), (b) W18O49/-NH2UiO-66 (Copyrighted from Ref. [30] with a license), (c) MIL-125@ZIF-67 (Copyrighted from Ref. [32] with a license), (d) Co2N0.67@ZIF-67 (Copyrighted from Ref. [33] with a license), and (e) CMS@MOFs (Copyrighted from Ref. [35] with a license)In response to the mass transfer limitations caused by the microporous structure of MOFs, Jiang et al. have developed a Co2N0.67@ZIF-67 composite by low-temperature pyrolysis of ZIF-67, where Co2N0.67 nanoparticles with photoconversion properties were generated in situ while maintaining the integrity of the main framework (Fig. 3d)[33]. The material optimizes substrate mass transfer efficiency through open pores formed by partial ligand dissociation, resulting in a more than 4-fold yield increase in the cycloaddition reaction of epichlorohydrin compared to the original ZIF-67 (22%). Wang et al. developed a cobalt nanoparticle-encapsulated doped with N hierarchical mesoporous carbon/titanium dioxide/MXene composite through controlled pyrolysis of a multilayered ZIF/MXene precursor. The engineered mesoporous architecture enhanced mass diffusion and photogenerated carrier separation while inducing internal light reflection to boost photothermal conversion efficiency, achieving a 96% cyclic carbonate yield under full-spectrum irradiation[34]. Substrate engineering provides an innovative solution to the problem of agglomeration in MOF photocatalysis. The CMS@MOFs composites exhibited enhanced mass transfer efficiency through a 3D interconnected pore network and hierarchical porosity design. The carbonized melamine sponge (CMS) substrate, with its high specific surface area, uniformly disperses MIL-88-NH2/ZIF-8-NH2 particles while preventing active site clogging. The hierarchical pore structure shortens reactant migration paths and reduces diffusion resistance (Fig. 3e)[35].
Therefore, constructing composite materials represents a powerful strategy to integrate the advantages of various components and achieve synergistic performance enhancement. It can simultaneously address multiple challenges faced by MOFs in photocatalysis, such as light harvesting, charge separation, and mass transfer. The current core challenge and future direction lie in the precise control of interfacial properties: how to rationally design and construct an ideal interface with strengthened chemical bonding and efficient carrier transport channels, rather than simple physical mixing.
2.3 Dopant incorporation
Dopant incorporation, involving the incorporation of metal or non-metal species into MOFs′ frameworks, has emerged as a powerful strategy to precisely tailor their electronic structure, light-harvesting properties, and catalytic activity.
By incorporating highly charged ions[51] or transition metals[52], the band structure and charge carrier dynamics of MOFs can be effectively modulated. Zhai et al. designed a Bi3+-functionalized porphyrinic MOF PCN-224(Zr), boosting cyclic carbonate yield from 31% (pristine PCN-224(Zr)) to 99% within 6 h due to optimized CO2 activation (Fig. 4a)[43]. Rare-earth elements further introduce unique 4f-orbital electron configurations, enabling multi-electron transfer processes, as demonstrated by Tb3+-exchanged DBP(Co), which attains a 99.3% yield in CO2 cycloaddition[44]. Beyond metal doping, single-atom incorporation offers atomic-level precision in active site engineering. Atomic-level doping creates well-defined active sites without compromising framework integrity. Ⅴ-doped indium-based MOF SNNU-97(In) enhances CO2 adsorption by 1.7 times and extends light absorption to 790 nm due to newly introduced electronic transitions (Fig. 4b)[45]. Besides, non-metal doping (e.g., boron in zirconium-based MOF UiO-67-B)[42] introduces distinct p-block electronic effects, enhancing Lewis acidity, charge separation, and chemisorption sites, leading to significantly accelerated reaction kinetics (Fig. 4c).
Figure 4
Dopant incorporation could effectively adjust the electronic structure of MOFs, enabling fine control over their light absorption range, carrier concentration, and separation efficiency, as well as surface acidity, without altering the main framework structure. However, the challenge of this strategy lies in achieving uniform distribution of doping elements within the framework and precise control of doping concentration to avoid the formation of electron recombination centers or phase separation.
3. Future directions and challenges for MOF-based photocatalysts
While significant progress has been achieved in developing MOF-based photocatalysts for CO2-to-cyclic carbonate conversion, critical challenges remain that demand the establishment of a synergistic framework integrating photon harvesting, charge transport, and surface activation.
For light absorption enhancement, breakthroughs beyond visible-light dependence require innovations at both molecular and mesoscopic levels. Bio-inspired designs mimicking chlorophyll′s broadband capture mechanisms could merge the bulk absorption advantages of narrow-bandgap semiconductors (e.g., black phosphorus, CuInS2) with the localized field enhancement of plasmonic materials (e.g., Au nanorods, Ag nanocubes), constructing hierarchical photon-harvesting networks. Simultaneously, the confinement effects of MOFs′ pores may enable precise regulation of quantum dot size distribution and spatial orientation, optimizing exciton dissociation pathways for efficient near-infrared photon conversion.
For the charge carrier′s management, cross-scale charge transport channels must be engineered: At the microscale, atomic layer deposition could tailor chemical bonding modes between MOFs and cocatalysts, designing low-barrier interfaces through orbital hybridization. At the mesoscale and macroscale, the foremost challenge currently is the precise modulation of interfacial dynamics between MOFs and other functional materials, such as semiconductors, carbon materials, and metal oxides, which critically influence photocatalytic performance. Existing methods for interfacial modulation lack the precision needed, leading to suboptimal charge transfer efficiency. Therefore, future endeavors should focus on developing advanced synthesis techniques like atomic layer deposition, exploring precise regulation methods such as molecular self-assembly, and refining interfacial structure design strategies. Furthermore, integrating in situ characterization with multiscale simulations will unravel the full dynamics of photogenerated carriers, from femtosecond-scale generation to millisecond-scale transport, establishing structure-activity models that correlate electronic states, interface density, and reaction pathways.
Beyond material design, transitioning MOF-based photocatalysts from laboratory prototypes to industrial-scale applications presents multifaceted challenges. Key considerations include (1) the scalable and cost-effective synthesis of these often complex materials, (2) the long-term chemical and structural stability under continuous irradiation and in flow reactor systems, (3) engineering efficient photoreactors that ensure optimal light distribution and photon efficiency throughout the catalyst bed, and (4) conducting thorough techno-economic analysis and life-cycle assessment to validate the environmental and economic viability of the process compared to conventional thermal catalysis. Addressing these engineering and economic hurdles is equally as important as improving the catalyst′s intrinsic activity to realize the practical potential of MOFs-driven photocatalytic CO2 valorization.
These insights will inform the rational design of next-generation adaptive photocatalysts. Ultimately, converging material genomics with reactor engineering will catalyze the transition from laboratory prototypes to industrial-scale photoreactor systems, achieving sunlight-driven CO2 valorization with atomic precision and macroscopic robustness.
-
-
[1]
KHAN M, AKMAL Z, TAYYAB M, MANSOOR S, LIU D N, YE Z W, ZHANG J L, WU S Q, WANG L Z. Integration of CO2 activation and photogenerated electron accumulation at Ti site via dual-tandem electric fields in BiOBr-MIL-125 heterojunction for boosting CO2 photoreduction[J]. Appl. Catal. B-Environ., 2025, 370: 125165.
-
[2]
WANG D D, XU M Y, LIN Z X, WU J H, YANG W T, LI H J, SU Z M. One-pot synthesis of MIL-68(In)-derived CdIn2S4/In2S3 tubular heterojunction for highly selective CO2 photoreduction[J]. Rare Metals, 2025, 44(6): 3956-3969.
-
[3]
LIU H, CHANG X L, YAN T, PAN W G. Research progress in catalytic conversion of CO2 and epoxides based on ionic liquids to cyclic carbonates[J]. Nano Energy, 2025, 135: 110596.
-
[4]
HUANG K, ZHANG J Y, LIU F, DAI S. Synthesis of porous polymeric catalysts for the conversion of carbon dioxide[J]. ACS Catal., 2018, 8: 9079-9102.
-
[5]
ZHAO C C, SU X F, LI R H, YAN L K, SU Z M. Insight into the mechanism of CO2 chemical fixation into epoxides by Windmill-shaped polyoxovanadate and n-Bu4NX (X=Br, I)[J]. Inorg. Chem., 2024, 63(30): 14032-14039.
-
[6]
CAI M L, DAI S Y, XUAN J, MO Y M. Bromide-mediated membraneless electrosynthesis of ethylene carbonate from CO2 and ethylene[J]. Nat. Commun., 2025, 16(1): 3285.
-
[7]
CHENG R L, WANG A H, SANG S X, LIANG H G, LIU S Q, TSIAKARAS P. Photocatalytic CO2 cycloaddition over highly efficient W18O49-based composites: An economic and ecofriendly choice[J]. Chem. Eng. J., 2023, 466: 142982.
-
[8]
LI D D, KASSYMOVA M, CAI X C, ZANG S Q, JIANG H L. Photocatalytic CO2 reduction over metal-organic framework-based materials[J]. Coordin. Chem. Rev., 2020, 412: 213262. doi: 10.1016/j.ccr.2020.213262
-
[9]
PRAJAPATI P K, KUMAR A, JAIN S L. First photocatalytic synthesis of cyclic carbonates from CO2 and epoxides using CoPc/TiO2 hybrid under mild conditions[J]. ACS Sustain. Chem. Eng, 2018, 6: 7799-7809. doi: 10.1021/acssuschemeng.8b00755
-
[10]
ZHANG H G, SI S H, ZHAI G Y, LI Y J, LIU Y Y, CHENG H F, WANG Z Y, WANG P, ZHENG Z K, DAI Y, LIU T X, HUANG B B. The long-distance charge transfer process in ferrocene-based MOFs with FeO6 clusters boosts photocatalytic CO2 chemical fixation[J]. Appl. Catal. B-Environ., 2023, 337: 122909.
-
[11]
YANG Q H, PENG H T, ZHANG Q J, QIAN X, CHEN X, TANG X, DAI S, ZHAO J J, JIANG K, YANG Q, SUN J, ZHANG L J, ZHANG N, GAO H L, LU Z Y, CHEN L. Atomically dispersed high-density Al-N4 sites in porous carbon for efficient photodriven CO2 cycloaddition[J]. Adv. Mater., 2021, 33(45): 2103186.
-
[12]
CHENG R L, NIU X Y, LI H, LIANG H G, TSIAKARAS P. Oxygen vacancy-rich defective tungsten oxide (WO3-x) modified by Prussian blue for efficient photocatalytic carbon dioxide conversion and tetracycline degradation[J]. J. Colloid Interface Sci., 2025, 683: 807-816. doi: 10.1016/j.jcis.2024.12.146
-
[13]
LIN Y P, LI L, SHI Z, ZHANG L S, LI K, CHEN J M, WANG H, LEE J M. Catalysis with two-dimensional metal-organic frameworks: Synthesis, characterization, and modulation[J]. Small, 2024, 20(24): 2309841. doi: 10.1002/smll.202309841
-
[14]
ZHANG H, LI C, LANG F F, LI M, LIU H Y, ZHONG D C, QIN J S, DI Z Y, WANG D H, ZENG L, PANG J D, BU X H. Precisely tuning band gaps of hexabenzocoronene-based MOFs toward enhanced photocatalysis[J]. Angew. Chem.-Int. Edit., 2025, 64(6): e202418017. doi: 10.1002/anie.202418017
-
[15]
SHEN Y, PAN T, WANG L, REN Z, ZHANG W N, HUO F W. Programmable logic in metal-organic frameworks for catalysis[J]. Adv. Mater., 2021, 33(46): 2007442. doi: 10.1002/adma.202007442
-
[16]
CUI W G, ZHANG G Y, HU T L, BU X H. Metal-organic framework-based heterogeneous catalysts for the conversion of C1 chemistry: CO, CO2 and CH4[J]. Coordin. Chem. Rev., 2019, 387: 79-120. doi: 10.1016/j.ccr.2019.02.001
-
[17]
LIANG J X, JIANG X, ZHANG X R, YU H, SHI J J, WANG M. Co-porphyrin-based metal-organic framework for light-driven efficient green conversion of CO2 and epoxides[J]. Chem. Eng. J., 2024, 499: 156428. doi: 10.1016/j.cej.2024.156428
-
[18]
KEGERE J, ALNEYADI S S, PAZ A P, SIDDIG L A, ALBLOOSHI A, ALNAQBI M A, ALZAMLY A, GREISH Y E. Titanium metal-organic frameworks for photocatalytic CO2 conversion through a cycloaddition reaction[J]. Nanoscale Adv., 2024, 6: 4804-4813. doi: 10.1039/D4NA00535J
-
[19]
HUANG Z W, HU K Q, MEI L, KONG X H, YU J P, LIU K, ZENG L W, CHAI Z F, SHI W Q. A mixed-ligand strategy regulates thorium-based MOFs[J]. Dalton Trans., 2020, 49(4): 983-987. doi: 10.1039/C9DT04158C
-
[20]
WU Y L, GAO L, ZHOU X C, YU X, MENG Y R, ZUO J L, SU J, YUAN S. Designing photothermal catalytic systems in multi-component MOFs for enhanced conversion of carbon dioxide[J]. Chem. Commun., 2024, 60(72): 9825-9828. doi: 10.1039/D4CC03203A
-
[21]
SHARMA N, DHANKHAR S S, NAGARAJA C M. A Mn(Ⅱ)-porphyrin based metal-organic framework (MOF) for visible-light-assisted cycloaddition of carbon dioxide with epoxides[J]. Microporous Mesoporous Mat., 2019, 280: 372-378. doi: 10.1016/j.micromeso.2019.02.026
-
[22]
SHI Q, CHEN M H, XIONG J, LI T, FENG Y Q, ZHANG B. Porphyrin-based two-dimensional metal-organic framework nanosheets for efficient photocatalytic CO2 transformation[J]. Chem. Eng. J., 2024, 481: 148301. doi: 10.1016/j.cej.2023.148301
-
[23]
ZHANG H G, ZHAI G Y, LEI L F, ZHANG C Y, LIU Y Y, WANG Z Y, CHENG H F, ZHENG Z K, WANG P, DAI Y, HUANG B B. Photo-induced photo-thermal synergy effect leading to efficient CO2 cycloaddition with epoxide over a Fe-based metal organic framework[J]. J. Colloid Interface Sci., 2022, 625: 33-41. doi: 10.1016/j.jcis.2022.05.146
-
[24]
ZHOU X L, ZHANG H G, CHENG H F, WANG Z Y, WANG P, ZHENG Z K, DAI Y, XING D N, LIU Y Y, HUANG B B. Enhanced cycloaddition between CO2 and epoxide over a bismuth-based metal organic framework due to a synergistic photocatalytic and photothermal effect[J]. J. Colloid Interface Sci., 2024, 658: 805-814. doi: 10.1016/j.jcis.2023.12.112
-
[25]
LI L, LIU W X, SHI T, SHANG S, ZHANG X D, WANG H, TIAN Z Q, CHEN L, XIE Y. Photoexcited single-electron transfer for efficient green synthesis of cyclic carbonate from CO2[J]. ACS Mater. Lett., 2023, 5(4): 1219-1226. doi: 10.1021/acsmaterialslett.3c00069
-
[26]
SIDDIG L A, ALZARD R H, NGUYEN H, GÖB C R, ALNAQBI M A, ALZAMLY A. Hexagonal layer manganese metal-organic framework for photocatalytic CO2 cycloaddition reaction[J]. ACS Omega, 2022, 7(11): 9958-9963. doi: 10.1021/acsomega.2c00663
-
[27]
SIDDIG L A, ALZARD R H, ABDELHAMID A S, RAMACHANDRAN T, NGUYEN H, PAZ A P, ALZAMLY A. Cobalt hydrogen-bonded organic framework as a visible light-driven photocatalyst for CO2 cycloaddition reaction[J]. Inorg. Chem., 2023, 62(38): 15550-15564. doi: 10.1021/acs.inorgchem.3c02051
-
[28]
LIU L F, ZHANG J L, CHENG X Y, XU M Z, KANG X C, WAN Q, HAN B X, WU N N, ZHENG L R, MA C Y. Amorphous NH2-MIL-68 as an efficient electro- and photo-catalyst for CO2 conversion reactions[J]. Nano Res., 2023, 16(1): 181-188. doi: 10.1007/s12274-022-4664-0
-
[29]
WU Z, GE Q M, CONG H, JIANG N, ZHAO W F, YANG S. Ligand-regulated oxygen vacancies in Ti-MOFs for visible-light-driven CO2 cycloaddition to cyclic carbonates[J]. Sep. Purif. Technol., 2025, 366: 132831. doi: 10.1016/j.seppur.2025.132831
-
[30]
程若霖, 王浩然, 任静, 马莹莹, 梁华根. W18O49/NH2-UiO-66复合催化剂的高效光催化CO2环加成[J]. 无机化学学报, 2024,40,(3): 523-532. CHENG R L, WANG H R, REN J, MA Y Y, LIANG H G. Efficient photocatalytic CO2 cycloaddition over W18O49/NH2-UiO-66 composite catalyst[J]. Chinese J. Inorg. Chem., 2024, 40(3): 523-532.
-
[31]
AHMED S H, BAKIRO M, ALZAMLY A. Photocatalytic activities of FeNbO4/NH2-MIL-125(Ti) composites toward the cycloaddition of CO2 to propylene oxide[J]. Molecules, 2021, 26(6): 1693. doi: 10.3390/molecules26061693
-
[32]
SHEN Q Y, CHEN W R, WANG M, JIN X X, ZHANG L X, SHI J L. A MOF@MOF S-scheme heterojunction with Lewis acid-base sites synergistically boosts cocatalyst-free CO2 cycloaddition[J]. ChemSusChem, 2025, 18(2): .
-
[33]
JIANG L J, WU D Q, HUANG Z S, CHEN F F, CHEN K, IBRAGIMOV A B, GAO J K. In situ pyrolysis of ZIF-67 to construct Co2N0.67@ZIF-67 for photocatalytic CO2 cycloaddition reaction[J]. Inorg. Chem., 2024, 63(31): 14761-14769. doi: 10.1021/acs.inorgchem.4c02504
-
[34]
WANG Y, DING M L, RONG W, KONG S Y, YAO J F. In situ growth of Fe-doped zeolitic imidazolate framework on MXene for boosting photodriven CO2 cycloaddition[J]. Sep. Purif. Technol., 2024, 345: 127399. doi: 10.1016/j.seppur.2024.127399
-
[35]
DUAN C Y, XIE Y M, DING M L, FENG Y, YAO J F. Design of carbonized melamine sponge@MOFs composites bearing diverse acid-base properties for boosting thermal and solar-driven CO2 cycloaddition[J]. J. CO2 Util., 2022, 64: 102158. doi: 10.1016/j.jcou.2022.102158
-
[36]
GONG L, SUN J, LIU Y S, YANG G C. Photoinduced synergistic catalysis on Zn single-atom-loaded hierarchical porous carbon for highly efficient CO2 cycloaddition conversion[J]. J. Mater. Chem. A, 2021, 9(38): 21689-21694. doi: 10.1039/D1TA06159C
-
[37]
WANG Y, DING M L, FU X T, YAO J F. High-efficiency photodriven coupling of CO2 and various epoxides via a multi-shelled hollow ZIF/MXene derived composite with low activation energy[J]. Inorg. Chem. Front., 2025, 12(3): 1114-1124. doi: 10.1039/D4QI02269F
-
[38]
ZHU H Y, SHEN Q Y, YUAN Y Y, GAO H, ZHOU S, YANG F L, SUN L M, WANG X J, YI J J, HAN X G. Engineering the sulfide semiconductor/photoinactive-MOF heterostructure with a hollow cuboctahedral structure to enhance photocatalytic CO2-epoxide-cycloaddition efficiency[J]. Inorg. Chem., 2024, 63(9): 4078-4085. doi: 10.1021/acs.inorgchem.3c03683
-
[39]
YANG G P, WU D, LU X M, TANG Y, GAO F, WANG Y Y. Light-assisted CO2 cycloaddition over a nanochannel cadmium-organic framework loaded with silver nanoparticles[J]. ACS Appl. Nano Mater., 2023, 6(7): 6197-6207. doi: 10.1021/acsanm.3c00499
-
[40]
GUO Y C, FENG L, WU C C, WANG X M, ZHANG X. Confined pyrolysis transformation of ZIF-8 to hierarchically ordered porous Zn-N-C nanoreactor for efficient CO2 photoconversion under mild conditions[J]. J. Catal., 2020, 390: 213-223. doi: 10.1016/j.jcat.2020.07.037
-
[41]
FANG Z, DENG Z, WAN X Y, LI Z Y, MA X, HUSSAIN S, YE Z Z, PENG X S. Keggin-type polyoxometalates molecularly loaded in Zr-ferrocene metal organic framework nanosheets for solar-driven CO2 cycloaddition[J]. Appl. Catal. B-Environ., 2021, 296: 120329. doi: 10.1016/j.apcatb.2021.120329
-
[42]
LI Y J, ZHAI G Y, LIU Y Y, WANG Z Y, WANG P, ZHENG Z K, CHENG H F, DAI Y, HUANG B B. Synergistic effect between boron containing metal-organic frameworks and light leading to enhanced CO2 cycloaddition with epoxides[J]. Chem. Eng. J., 2022, 437(1): 135363.
-
[43]
ZHAI G Y, LIU Y Y, LEI L F, WANG J J, WANG Z Y, ZHENG Z K, WANG P, CHENG H F, DAI Y, HUANG B B. Light-promoted CO2 conversion from epoxides to cyclic carbonates at ambient conditions over a bi-based metal-organic framework[J]. ACS Catal., 2021, 11(4): 1988-1994. doi: 10.1021/acscatal.0c05145
-
[44]
SHI Q, XU Y H, CHEN M H, ZHAO P, DENG W B, XIONG J, GUO K, FENG Y Q, ZHANG B. A versatile synthetic strategy towards rare earth based metal-organic frameworks[J]. Sci. China Chem., 2025, 68(4): 1362-1371. doi: 10.1007/s11426-024-2322-9
-
[45]
FAN S C, CHEN S Q, WANG J W, LI Y P, ZHANG P, WANG Y, YUAN W Y, ZHAI Q G. Precise introduction of single vanadium site into indium-organic framework for CO2 capture and photocatalytic fixation[J]. Inorg. Chem., 2022, 61: 14131-14139. doi: 10.1021/acs.inorgchem.2c02250
-
[46]
ZHAI G Y, LIU Y Y, MAO Y Y, ZHANG H G, LIN L T, LI Y J, WANG Z Y, CHENG H F, WANG P, ZHENG Z K, DAI Y, HUANG B B. Improved photocatalytic CO2 and epoxides cycloaddition via the synergistic effect of Lewis acidity and charge separation over Zn modified UiO-bpydc[J]. Appl. Catal. B-Environ., 2022, 301: 120793. doi: 10.1016/j.apcatb.2021.120793
-
[47]
HUANG Z W, HU K Q, MEI L, WANG C Z, CHEN Y M, WU W S, CHAI Z F, SHI W Q. Potassium ions induced framework interpenetration for enhancing the stability of uranium-based porphyrin MOF with visible-light-driven photocatalytic activity[J]. Inorg. Chem., 2021, 60(2): 652-660.
-
[48]
FANG Z B, LIU T T, LIU J X, JIN S Y, WU X P, GONG X Q, WANG K C, YIN Q, LIU T F, CAO R, ZHOU H C. Boosting interfacial charge-transfer kinetics for efficient overall CO2 photoreduction via rational design of coordination spheres on metal-organic frameworks[J]. J. Am. Chem. Soc., 2020, 142(28): 12515-12523. doi: 10.1021/jacs.0c05530
-
[49]
DAS R, MANNA S S, PATHAK B, NAGARAJA C M. Strategic design of Mg-centered porphyrin metal-organic framework for efficient visible light-promoted fixation of CO2 under ambient conditions: Combined experimental and theoretical investigation[J]. ACS Appl. Mater. Interfaces, 2022, 14(29): 33285-33296. doi: 10.1021/acsami.2c07969
-
[50]
CUI W G, HU T L. Incorporation of active metal species in crystalline porous materials for highly efficient synergetic catalysis[J]. Small, 2021, 17: 2003971. doi: 10.1002/smll.202003971
-
[51]
LIU C, CHEN H L, CHEN Q, BI J H, YU J C, WU L. Active site modulation in UiO-66(Ce) MOFs by Al3+ doping for boosting photocatalysis[J]. J. Catal., 2024, 437: 115670. doi: 10.1016/j.jcat.2024.115670
-
[52]
ZHANG Y Y, HUANG R T, FANG Y, WANG J C, YUAN Z J, CHEN X W, ZHU W J, CAI Y, SHI X Y. Modulation of Fe-MOF via second-transition metal ion doping (Ti, Mn, Zn, Cu) for efficient visible-light driven CO2 reduction to CH4[J]. Sep. Purif. Technol., 2024, 336: 126164. doi: 10.1016/j.seppur.2023.126164
-
[1]
-
Figure 3 Mechanism diagrams of photocatalytic cyclic carbonates production: (a) PMo12@Zr-Fc (Copyrighted from Ref. [41] with a license), (b) W18O49/-NH2UiO-66 (Copyrighted from Ref. [30] with a license), (c) MIL-125@ZIF-67 (Copyrighted from Ref. [32] with a license), (d) Co2N0.67@ZIF-67 (Copyrighted from Ref. [33] with a license), and (e) CMS@MOFs (Copyrighted from Ref. [35] with a license)
Table 1. Performance comparison of MOFs for photocatalytic CO2 cycloaddition
Catalyst Light source Time /h Reaction substrate Conversion /% Yield /(mmol·g-1·h-1) Regulatory strategy Ref. Fe-FcDC 300 W xenon lamp 4 Epichlorohydrin 7 44 Ligand engineering [10] Co-PMOF 3 300 W xenon lamp 6 Epichlorohydrin > 99 61 Ligand engineering [17] MOF-997 300 W xenon lamp 24 Styrene oxide > 99 2 Ligand engineering [18] Th-IHEP-5 45 W fluorescent light 48 Styrene oxide 71 0.6 Ligand engineering [19] Zr-TTF-L1 300 W xenon lamp 8 Propylene oxide 45 14 Ligand engineering [20] MOF 1 (Mn) LED 18×3W 24 Epichlorohydrin 100 17 Ligand engineering [21] Fe-DBP(Co) 1000 W xenon lamp 12 Epichlorohydrin 97 103 Ligand engineering [22] FeTPyP 300 W xenon lamp 5 Styrene oxide 30 106 Ligand engineering [23] Bi-HHTP 300 W xenon lamp 4 Styrene oxide 34 112 Ligand engineering [24] Mg-MOF-74 400 mW·cm-2 12 1-Bromo-2,3-epoxypropane 97 5 Ligand engineering [25] UAEU-50 400 W halogen lamp 24 Styrene oxide 90 6 Ligand engineering [26] Co-HOF 400 W halogen lamp 24 Styrene oxide > 99 6 Ligand engineering [27] NH2-MIL-68 300 W xenon lamp 12 Styrene oxide 94 56 Ligand engineering [28] Ti-H2TPDC 8 W LED light 4 Epichlorohydrin 94 59 Ligand engineering [29] PB-modified WO3-x 300 W xenon lamp 8 Styrene oxide 94 171 Composite construction [12] W18O49/NH2-UiO-66 300 W xenon lamp 4 Styrene oxide 82 58 Composite construction [30] FeNbO4/NH2-MIL-125(Ti) 500 W halogen lamp 72 Propylene oxide 52 0.2 Composite construction [31] MIL-125@ZIF-67 3 W blue LED 36 1-Bromo-2,3-epoxypropane 99 0.3 Composite construction [32] Co2N0.67@ZIF-67 350 mW·cm-2 3 Epichlorohydrin 95 3 Composite construction [33] Zn20Fe1-ZIF/MXene 350 mW·cm-2 6 Epichlorohydrin 96 16 Composite construction [34] CMS@MIL-88-NH2 300 W xenon lamp 6 1, 2-Epoxybutane 72 24 Composite construction [35] Zn SA-NC 300 mW·cm-2 16 Epichlorohydrin 99 0.5 Composite construction [36] CoNHPC/TM-8 350 mW·cm-2 12 Styrene oxide 99 8 Composite construction [37] CuS@HKUST-1 Blue light 18 2-Chloromethyl-oxriane 98 11 Composite construction [38] Ag0.1/NWU-Cd3a 300 W xenon lamp 6 Styrene oxide 94 22 Composite construction [39] Zn-N-HOPCPs 350 mW·cm-2 10 Propylene oxide 95 29 Composite construction [40] PMo12@Zr-Fc 500 mW·cm-2 8 Epichlorohydrin 96 298 Composite construction [41] UiO-67-B 300 W xenon lamp 6 Propylene epoxide 6 45 Dopant incorporation [42] Bi-PCN-224 300 W xenon lamp 6 Propylene epoxide 99 5 Dopant incorporation [43] Tb-DBP(Co) 300 W xenon lamp 10 Epichlorohydrin > 99 127 Dopant incorporation [44] SNNU-97-InV 300 W xenon lamp 24 2-Bromooxirane 97 14 Dopant incorporation [45] Zn-UiO-bpydc 300 W xenon lamp 9 Propylene oxide 75 0.3 Dopant incorporation [46] IHEP-9 45 W compact fluorescent lamp 12 Styrene oxide 99 2 Dopant incorporation [47] -
扫一扫看文章
计量
- PDF下载量: 0
- 文章访问数: 22
- HTML全文浏览量: 3

下载:
下载:
下载: