Skip to main page content
U.S. flag

An official website of the United States government

Dot gov

The .gov means it’s official.
Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you’re on a federal government site.

Https

The site is secure.
The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.

Access keys NCBI Homepage MyNCBI Homepage Main Content Main Navigation
Review
. 2023 May 10;123(9):5347-5420.
doi: 10.1021/acs.chemrev.2c00879. Epub 2023 Apr 12.

Bioinspired Framework Catalysts: From Enzyme Immobilization to Biomimetic Catalysis

Affiliations
Review

Bioinspired Framework Catalysts: From Enzyme Immobilization to Biomimetic Catalysis

Kun-Yu Wang et al. Chem Rev. .

Abstract

Enzymatic catalysis has fueled considerable interest from chemists due to its high efficiency and selectivity. However, the structural complexity and vulnerability hamper the application potentials of enzymes. Driven by the practical demand for chemical conversion, there is a long-sought quest for bioinspired catalysts reproducing and even surpassing the functions of natural enzymes. As nanoporous materials with high surface areas and crystallinity, metal-organic frameworks (MOFs) represent an exquisite case of how natural enzymes and their active sites are integrated into porous solids, affording bioinspired heterogeneous catalysts with superior stability and customizable structures. In this review, we comprehensively summarize the advances of bioinspired MOFs for catalysis, discuss the design principle of various MOF-based catalysts, such as MOF-enzyme composites and MOFs embedded with active sites, and explore the utility of these catalysts in different reactions. The advantages of MOFs as enzyme mimetics are also highlighted, including confinement, templating effects, and functionality, in comparison with homogeneous supramolecular catalysts. A perspective is provided to discuss potential solutions addressing current challenges in MOF catalysis.

PubMed Disclaimer

Conflict of interest statement

The authors declare no competing financial interest.

Figures

Figure 1
Figure 1
Overview of strategies to synthesize bioinspired MOF catalysts. (top) Enzymes are incorporated into MOFs to afford biocomposites. (bottom) Model compounds emulating enzyme’s active sites can be introduced into MOFs through guest encapsulation, metal node functionalization, and organic ligand functionalization.
Figure 2
Figure 2
Overview of strategies to prepare MOF–enzyme composites, including surface attachment, encapsulation, covalent linkage, and coprecipitation. Surface attachment directly anchors enzymes to MOFs’ surfaces via noncovalent interactions, including hydrophobic interactions, van der Waals forces, and electrostatic forces. Encapsulation indicates entirely absorbing the enzymes into the pores of MOFs and establishing interactions within the interior environment. Covalent linkage utilizes the functional groups on both MOFs and enzymes to form covalent bonding. Coprecipitation refers to mixing up enzymes and the reactants of MOFs in the homogeneous phase, embedding the enzymes in the instantaneously formed pores.
Figure 3
Figure 3
Schematic illustration of enzyme immobilization methods in MOFs. (a) PCN-222 as the supporter for immobilization of GOx by electrostatic interaction. Reproduced from ref (100). Copyright 2019 American Chemical Society. (b) The coordinative bond between the imidazole group from MO act as Lewis base and coordinatively unsaturated metal sites (CUS) acting as Lewis acid in immobilization. Reproduced from ref (106). Copyright 2017 American Chemical Society. (c) Enzyme encapsulation where formate dehydrogenase infiltrates into the pores of NU-1006. Reproduced with permission from ref (141). Copyright 2019 John Wiley and Sons. (d) Covalent linkage via N-hydroxysuccinimide to immobilize GOx on NH2-MIL-53(Al). Reproduced with permission from ref (112). Copyright 2016 Royal Society of Chemistry. (e) Illustration showing coprecipitation and biomimetic mineralization via a one-pot synthesis to immobilize urease in ZIF-8. Reproduced with permission from ref (153). Copyright 2016 Royal Society of Chemistry.
Figure 4
Figure 4
Encapsulation of cutinase into the mesopores of NU-1000. Reproduced with permission from ref (125). Copyright 2016 Elsevier.
Figure 5
Figure 5
One-pot synthesis of BCL@MTV-ZIFs, in which the closed-lid/open-lid conformations of BCL were regulated via MTV-ZIFs. Reproduced with permission from ref (138). Copyright 2021 American Chemical Society.
Figure 6
Figure 6
Utilization of polypeptide to boost the stability of ZIFs toward acid treatment. Reproduced with permission from ref (130). Copyright 2021 John Wiley and Sons.
Figure 7
Figure 7
Hydrolysis reaction for chemical warfare agent degradation using enzyme OPAA encapsulated within NU-1003. Reproduced with permission from ref (125). Copyright 2016 Elsevier.
Figure 8
Figure 8
Glucose oxidase immobilized on 2D MOF to conduct an oxidation reaction. Oxygen is used as an oxidant to oxidize glucose to generate radicals to kill bacteria. Reproduced with permission from ref (178). Copyright 2019 American Chemical Society.
Figure 9
Figure 9
Catalase immobilized on MAF-7 and ZIF-90. The catalytic performances of catalase are presented to demonstrate residue H2O2 concentration decrease with a dependence on the time. Reproduced with permission from ref (50). Copyright 2019 American Chemical Society.
Figure 10
Figure 10
FateDH immobilized in ZIF-8 carrying out carbon sequestration under light. Reproduced with permission from ref (188). Copyright 2021 Elsevier.
Figure 11
Figure 11
Two enzymes, namely Gox and HRP, immobilized in a MOF that together perform the oxidation of glucose and electron transport to water. Meanwhile, 3,3′,5,5′-tetramethylbenzidine (TMB) is oxidized to oxTMB. Reproduced with permission from ref (197). Copyright 2015 Royal Society of Chemistry.
Figure 12
Figure 12
Building block design of MOFs emulating the enzymatic active sites. Model compounds are integrated into MOFs as metal clusters and organic ligands to reproduce functions of enzymes. Two representative metal clusters are Zr6 cluster emulating phosphotriesterase and Zn cluster emulating carbonic anhydrase. Active sites, such as urea, diiron, and porphyrin, can be embedded onto the organic linkers.
Figure 13
Figure 13
Representative MOFs with open metal sites. (a) MOF-74 containing one-dimensional channels and M2O2 (CO2)2 metal-oxo chains. (b) HKUST-1 based on Cu2 (CO2)4 paddle-wheel clusters. (c) MIL-100 and MIL-101 consisting of M3O (CO2)6 clusters.
Figure 14
Figure 14
Pentapeptide cleavage using a multivariate MOF with enzyme-like structural complexity. Reproduced with permission from ref (282). Copyright 2016 American Chemical Society.
Figure 15
Figure 15
Long-range ordered arrangement of the urea/squaramide groups in frameworks helps to avoid the self-quenching of the active sites due to the oligomer formation for free urea/squaramide-based small molecules.
Figure 16
Figure 16
Structures and design principles for reported linkers used for constructing urea-based MOFs. The reported urea-based ligands consist of carboxylate groups or pyridyl groups. Urea groups can be incorporated into the backbone and substituent of the ligands, affording ligands with varied connectivities and configurations.
Figure 17
Figure 17
Enhanced catalytic activity of urea-based MOFs with the addition of Lewis acid. Reproduced with permission from ref (318). Copyright 2016 American Chemical Society.
Figure 18
Figure 18
Summary of nucleophilic substitution reactions catalyzed by urea- and/or squaramide-based MOFs.
Figure 19
Figure 19
Summary of reported linkers used for constructing squaramide-based MOFs.
Figure 20
Figure 20
Synthesis of multivariate squaramide-based UiO-67 and its catalytic activity toward Friedel–Crafts reaction. Reproduced with permission from ref (66). Copyright 2015 American Chemical Society.
Figure 21
Figure 21
Representative MOFs based on Zr6 clusters and organic linkers with varied connectivities, including ditopic, tritopic, tetratopic, hexatopic linkers, and their mixtures.
Figure 22
Figure 22
Structural illustrations of phosphotriesterase’s active site and a Zr6-based MOF named UiO-66. Reproduced with permission from ref (63). Copyright 2014 John Wiley and Sons.
Figure 23
Figure 23
High-throughput screening of Zr6-based MOFs for DMNP degradation, in which the catalytic activity at pH 8 and 10 were compared. Reproduced with permission from ref (410). Copyright 2018 Royal Society of Chemistry.
Figure 24
Figure 24
Hydrolysis of nerve agent simulators containing phosphonate linkages, which is catalyzed by phosphotriesterase inspired MOFs with Zr6O4 (OH4) clusters.
Figure 25
Figure 25
Fabrication of SILK@UiO-66@LiOtBu for CWA degradation. Reproduced with permission from ref (424). Copyright 2015 John Wiley and Sons.
Figure 26
Figure 26
Ligand design of dehydrogenase-mimicking MOFs. (a) Active sites of formate dehydrogenase (FDH). (b) Crystal structure of a dehydrogenase-mimicking MOF as an electrochemical glucose sensor. Reproduced with permission from ref (218). Copyright 2020 American Chemical Society.
Figure 27
Figure 27
Preparation of MOF-derived carbon with neighboring Fe and Ni single-atoms, which was applied in CO2 electroreduction. Reproduced with permission from ref (470). Copyright 2021 American Chemical Society.
Figure 28
Figure 28
Structural illustrations and CO2 capture mechanism of carbonic anhydrase. Reproduced with permission from ref (472). Copyright 2018 Springer Nature.
Figure 29
Figure 29
Two triazolate-based MOFs as mimics of carbonic anhydrase. (a) Active site of carbonic anhydrase. (b) Structures of CFA-1 and MFU-4l with exposed Zn sites. Reproduced with permission from ref (476). Copyright 2020 Royal Society of Chemistry.
Figure 30
Figure 30
CO2 chemisorption on the Zn–OH site through intercluster hydrogen bonding interactions. Reproduced with permission from ref (221). Copyright 2018 American Chemical Society.
Figure 31
Figure 31
Subunits and cofactors of Mo nitrogenase. Reproduced with permission from ref (489). Copyright 1994 American Chemical Society.
Figure 32
Figure 32
Brief summary of catalysis reaction conducted by alternative nitrogenases. Reproduced with permission from ref (523). Copyright 2020 American Chemical Society.
Figure 33
Figure 33
Structural illustration of a MOF V2Cl2.8 (btdd) with accessible V site to bond nitrogen.
Figure 34
Figure 34
Synthesis of [Fe4S4]-based redox-active coordination polymers. Reproduced with permission from ref (228). Copyright 2019 American Chemical Society.
Figure 35
Figure 35
Illustration of hydrogenases and hydrogenase-mimicking ligands. (a) Active sites of [NiFe]–H2ases and [FeFe]–H2ases. (b) Ligand design of [FeFe]–H2ase-mimicking MOFs. Reproduced with permission from ref (561). Copyright 2014 American Chemical Society.
Figure 36
Figure 36
Three strategies to synthesize [FeFe]–H2ase-mimicking MOF. (a) Introducing [FeFe] ligand through ligand exchange to afford a multivariate MOF enabling photochemical hydrogen production. Reproduced with permission from ref (230). Copyright 2013 American Chemical Society. (b) Installing [FeFe] model compounds to accessible metal sites in a porphyrin MOF. Reproduced with permission from ref (232). Copyright 2014 Royal Society of Chemistry. (c) Integrating redox-active ligands and [FeFe]–H2ase active sites into a defective MOF through linker installation to mimic the electron transport chain. Reproduced with permission from ref (235). Copyright 2021 American Chemical Society.
Figure 37
Figure 37
Structural illustration of (a) Heme-b and (b) TPP. (c) Decomposition of Fe(TPP) SPh to produce μ-oxo-bridged dimers.
Figure 38
Figure 38
Crystal structure of PCN-222(Fe) with a csq network topology, enabling oxidizing pyrogallol by hydrogen peroxide. Reproduced with permission from ref (64). Copyright 2012 John Wiley and Sons.
Figure 39
Figure 39
Crystal structure and building blocks of PCN-600 for catalytic oxidation. Reproduced with permission from ref (65). Copyright 2014 American Chemical Society.
Figure 40
Figure 40
Preparation of mixed-metal-salen MOFs through postsynthetic linker exchange. The salen-based MOFs enable conducting multiple asymmetric catalysis efficiently. Reproduced with permission from ref (630). Copyright 2018 American Chemical Society.
Figure 41
Figure 41
Constructing chiral MOF@MOF composites from M-salen ligands for asymmetric epoxidation/ring-opening reactions. Reproduced with permission from ref (631). Copyright 2019 Royal Society of Chemistry.
Figure 42
Figure 42
Illustration of bimetallic MOF-919 (Fe–Cu) mimicking bifunctional oxidase–peroxidase catalytic activity. Reproduced with permission from ref (640). Copyright 2022 Royal Society of Chemistry.
Figure 43
Figure 43
Comparison between enzymatic catalysis, supramolecular catalysis, and framework catalysis.
Figure 44
Figure 44
Polymerization within MOFs with different channel sizes resulting in diverse products. Reproduced with permission from ref (645). Copyright 2007 John Wiley and Sons.
Figure 45
Figure 45
Fixation and [2+2+2] cyclotrimerization of substituted pyridines bearing unsaturated functional groups within MIL-88B (Fe). The locations of pyridine monomers are confirmed by SCXRD. Reproduced with permission from ref (647). Copyright 2015 Springer Nature.
Figure 46
Figure 46
(top) Illustration of the azide sites and the site-selective click reaction of a dialkyne within a Mn-based MOF. Bottom: The click reaction products of dialkynes with varied chain length. The site-selective click reaction involves the immobilization of azides and regeneration via alkylation with MeBr. Reproduced with permission from ref (648). Copyright 2018 American Chemical Society.
Figure 47
Figure 47
Quantitative hydrogenolysis to generate catalytic Ru@MOF-5 composites. Reproduced with permission from ref (651). Copyright 2005 John Wiley and Sons.
Figure 48
Figure 48
Overview of strategies to synthesizing hydrophobic MOFs or MOF-based composites. (a) Introducing hydrophobic modulator acid into the MOF. Reproduced with permission from ref (674). Copyright 2016 American Chemical Society. (b) Coating the MOF surface with poly(dimethylsiloxane) (PDMS). Reproduced with permission from ref (675). Copyright 2014 American Chemical Society. (c) Fabricating composites consisting of MOFs and highly fluorinated graphene (HFGO). Reproduced with permission from ref (679). Copyright 2016 John Wiley and Sons. (d) Functionalizing MOFs with aliphatic chains through covalent bonds. Reproduced with permission from ref (681). Copyright 2011 John Wiley and Sons.
Figure 49
Figure 49
Catalytic activity of a homochiral BINAP-MOF metalated with Rh and Ru. Reproduced with permission from ref (677). Copyright 2014 American Chemical Society.
Figure 50
Figure 50
Structural of Pd12L24 molecular assembly that enables increasing the gold concentration and improving catalytic efficiency. Reproduced with permission from ref (723). Copyright 2014 John Wiley and Sons.
Figure 51
Figure 51
Self-assembly of Co (II) phthalocyanine within the bio-MOF-1’s nanopores, involving metal cation exchange and cation-directed assembly. Reproduced with permission from ref (39). Copyright 2014 American Chemical Society.
Figure 52
Figure 52
MOF-templated stepwise synthesis of homo- (a) and heterobimetallic (b) supramolecular coordination compounds within the confined MOF channels. Step (i) indicates the incorporation of organic ligand with desired structural and coordination information. Step (ii) indicates postsynthetic metalation. Reproduced with permission from ref (732). Copyright 2019 American Chemical Society.
Figure 53
Figure 53
Installing sulfuric acid onto PCN-222 to generate a photoactive and acidic MOF, which can catalyze photocatalytic oxidation of dihydroartemisinic acid to artemisinin. Reproduced with permission from ref (733). Copyright 2019 American Chemical Society.
Figure 54
Figure 54
Construction of catalytic MOFs through postsynthetic linker installation. (a) PCN-700 features two types of missing-linker defects, which can accommodate carboxylate ligands with varied sizes. (b) Active sites were introducing to endow the MOF with catalytic activity. Reproduced with permission from ref (256). Copyright 2016 American Chemical Society.
Figure 55
Figure 55
Illustration of the pore environment in MUF-77 equipped with catalytic unit and modulator groups. The potential contacts between the aldo intermediate (orange) and the modulator groups (violet and red) are shown in the right. Reproduced with permission from ref (711). Copyright 2017 American Chemical Society.
Figure 56
Figure 56
Regio- and enantioselective hydroformylation of Rh-monophosphane complexes confined within cyclodextrins. Reproduced with permission from ref (746). Copyright 2014 John Wiley and Sons.
Figure 57
Figure 57
Regio- and enantioselective photodimerization confined within a cyclodextrin-based MOF. The substrates can form superstructures with cyclodextrin frameworks. Reproduced with permission from ref (749). Copyright 2021 American Chemical Society.
Figure 58
Figure 58
Endowing a gold nanocluster with water solubility and HRP-mimicking catalytic activity through bonding with CD-MOF-1. Reproduced with permission from ref (751). Copyright 2020 American Chemical Society.
Figure 59
Figure 59
Structural illustration of the calix[4]arene linker and the (4,6)-connected MOF. The two nonintersecting pores are depicted in the framework in red and orange. The N2 sorption isotherm at 77 K is displayed. Reproduced with permission from ref (764). Copyright 2018 John Wiley and Sons.
Figure 60
Figure 60
Dynamic self-assembly of a guest/host/metal–cation complex. The transition metal promoted the selective photoreactions within the cucurbiturils. Reproduced with permission from ref (773). Copyright 2011 John Wiley and Sons.
Figure 61
Figure 61
Structural illustration of the SMOF-1 and WD-POM@SMOF-1. Carbon, nitrogen, oxygen, and hydrogen are represented in gray, blue, red, and white, respectively. Reproduced with permission from ref (781). Copyright 2016 Springer Nature.
Figure 62
Figure 62
Self-assembly of cavitands to form a capsule to store steroids. Reproduced with permission from ref (783). Copyright 2004 American Chemical Society.
Figure 63
Figure 63
Distinct regioselectivity of Diels–Alder reaction confined within cage-shaped and bowl-shaped self-assembled container molecules. Reproduced with permission from ref (786). Copyright 2006 the American Association for the Advancement of Science.
Figure 64
Figure 64
Structural illustration of the [Ga4L6]12– cage and the catalytic mechanism for orthoformate hydrolysis. Reproduced with permission from ref (799). Copyright 2007 the American Association for the Advancement of Science.

Similar articles

Cited by

References

    1. Schlichting I.; Berendzen J.; Chu K.; Stock A. M.; Maves S. A.; Benson D. E.; Sweet R. M.; Ringe D.; Petsko G. A.; Sligar S. G. The Catalytic Pathway of Cytochrome P450cam at Atomic Resolution. Science 2000, 287, 1615–1622. 10.1126/science.287.5458.1615. - DOI - PubMed
    1. Dawson J. H.; Sono M. Cytochrome P-450 and Chloroperoxidase: Thiolate-Ligated Heme Enzymes. Spectroscopic Determination of Their Active-Site Structures and Mechanistic Implications of Thiolate Ligation. Chem. Rev. 1987, 87, 1255–1276. 10.1021/cr00081a015. - DOI
    1. Rittle J.; Green M. T. Cytochrome P450 Compound I: Capture, Characterization, and Ch Bond Activation Kinetics. Science 2010, 330, 933–937. 10.1126/science.1193478. - DOI - PubMed
    1. Groves J. T. Using Push to Get Pull. Nat. Chem. 2014, 6, 89–91. 10.1038/nchem.1855. - DOI - PMC - PubMed
    1. Breslow R.; Overman L. E. ″Artificial Enzyme″ Combining a Metal Catalytic Group and a Hydrophobic Binding Cavity. J. Am. Chem. Soc. 1970, 92, 1075–1077. 10.1021/ja00707a062. - DOI - PubMed

Publication types

Substances