Lulu
He
abc,
Le
Wang
ac,
Zhen
He
a,
Cheng Heng
Pang
*b,
Bencan
Tang
*b,
Aiguo
Wu
*ac and
Juan
Li
*ac
aNingbo Key Laboratory of Biomedical Imaging Probe Materials and Technology, Zhejiang International Cooperation Base of Biomedical Materials Technology and Application, Chinese Academy of Sciences (CAS) Key Laboratory of Magnetic Materials and Devices, Ningbo Cixi Institute of Biomedical Engineering, Zhejiang Engineering Research Center for Biomedical Materials, Ningbo Institute of Materials Technology and Engineering, Chinese Academy of Sciences, Ningbo, 315201, China. E-mail: aiguo@nimte.ac.cn; lij@nimte.ac.cn
bDepartment of Chemical and Environment Engineering, The University of Nottingham Ningbo China, Ningbo, 315100, China. E-mail: Chengheng.Pang@nottingham.edu.cn; Bencan.Tang@nottingham.edu.cn
cAdvanced Energy Science and Technology Guangdong Laboratory, Huizhou, 516000, China
First published on 19th December 2023
Covalent organic frameworks (COFs), a new and developing class of porous framework materials, are considered a type of promising carrier for the integration and delivery of bioactives, which have diverse fascinating merits, such as a large specific surface area, designable and specific porosity, stable and orderly framework structure, and various active sites. However, owing to the significant differences among bioactives (including drugs, proteins, nucleic acid, and exosomes), such as size, structure, and physicochemical properties, the interaction between COFs and bioactives also varies. In this review, we firstly summarize three strategies for the construction of single or hybrid COF-based matrices for the delivery of cargos, including encapsulation, covalent binding, and coordination bonding. Besides, their smart response release behaviors are also categorized. Subsequently, the applications of cargo@COF biocomposites in biomedicine are comprehensively summarized, including tumor therapy, central nervous system (CNS) modulation, biomarker analysis, bioimaging, and anti-bacterial therapy. Finally, the challenges and opportunities in this field are briefly discussed.
Wider impactCovalent organic frameworks (COFs) have received increasing attention owing to their potential applications in tumor therapy, central nervous system (CNS) modulation, biomarker analysis, bioimaging, and anti-bacterial therapy. More importantly, COFs with a high porosity, abundant active sites, and stimulus-responsive structures, usually function as host materials for the integration and delivery of bioactives in the above-mentioned practical applications. In this review, we summarize and discuss three types of synthetic strategies for the preparation of COF-based host–guest systems towards diverse bioactives, including small molecules, proteins, nucleic acids and exosomes. In addition, the recent advances in various types of COFs towards diverse biomedical applications are discussed in detail. This review also provides a complete overview of the current developments in COF-based nanosystems from the viewpoint of host–guest connection approach design and synthetic strategies, and their applicability in biomedical applications. We believe that the integration approaches of COFs and bioactives elucidated in this review will provide a significant basis for exploring next-generation COF-based integrated nanosystems. |
Here, bioactives not only refer to small molecules, such as therapeutic drugs, but also biomacromolecules, including proteins and nucleic acids, and even more complex vesicles, such as exosomes. Bioactives alone often fail to achieve a satisfactory therapeutic result owing to their inherent drawbacks. For example, conventional chemotherapeutic drugs are always limited by poor targeting ability and serious side effects.8 Furthermore, maintaining the activity of proteins and prolonging the half-life of nucleic acids in vivo are also some of the challenging issues.9 However, owing to the unique properties of COFs, they can overcome these difficulties, and thus have attracted extensive attention for the integration and delivery of bioactives.
The main advantages of COFs as carriers for the integration of bioactives are as follows: (1) diverse COF-based functional materials with different sizes, topologies, morphologies, and physical/chemical properties that can be designed and fabricated using available organic linkers and linkages, which can be summarized in one word: designability. (2) COFs with a high crystallinity tend to provide an extremely high specific surface area (over 1000 m2 g−1), uniform pore shape, ultrahigh porosity, and adjustable and wide pore size distribution (from micropore to mesopore), making them an excellent host matrix for encapsulating bioactives. (3) COFs possess a good biocompatibility and low cytotoxicity, which is attributed to their organic and metal-free nature. (4) Through the post-synthetic modification (PSM) strategy, bonding defect functionalization (BDF) strategy, or other surface decoration approach, target molecules can be covalently attached to the COF backbone by reacting with the active sites of its linkers, while retaining the original framework. (5) Smart stimuli-responsive COFs have been successfully designed for controlled and sustained drug release due to the physical and chemical difference in the microenvironment between healthy and diseased tissues.
This review focuses on the integration strategies for cargo@COF complexes and their diverse biomedical applications (Scheme 1). Thus far, several excellent reviews have discussed COFs and their wide applications, such as catalysis,10–12 gas storage and separation,13,14 vapor sorption,15 electronic and ionic conduction.16–18 However, only a few further presented the diverse bio-applications of COFs.19–21 Also, another limitation of existing reviews on COF-based carrier systems in the biomedical field is that almost all focused on cancer therapy, while ignoring other applications.22 Thus, in this review, we highlight the importance of COFs as host materials for the delivery of bioactives and their applications, not only in cancer therapy but also in other biomedical fields. Notably, the use of COFs as hosts for the integration and delivery of bioactives is still in its infancy, making it important to summarize and introduce this COF-based specific area, expanding the applications of COFs.
Scheme 1 Categorization of COFs as host materials for the integration and delivery of bioactives and their biomedical applications. |
Year | Carrier | Cargo | Composite strategy | Loading percentage | Applications | Ref. |
---|---|---|---|---|---|---|
2015 | PI-COF-4 | IBU/captopril/caffeine | Two-step encapsulation | ∼20 wt% for IBU | — | 55 |
PI-COF-5 | ||||||
2016 | PI-3-COF | 5-FU/captopril/IBU | Two-step encapsulation | 16 wt% | Drug delivery in vitro | 56 |
PI-2-COF | 30 wt% | |||||
2017 | PCTF | IBU | Two-step encapsulation | 19 wt% | — | 57 |
PCTF-Mn | 23 wt% | |||||
2017 | TpAPH | 5-FU | PSM | 12 wt% | Targeted drug delivery | 48 |
TpASH | ||||||
2018 | APTES-COF-1 | DOX | Two-step encapsulation | 9.71 ± 0.13 wt% | Chemotherapy | 58 |
2019 | TPB-DMTP-COF | Por | BDF | 0.091 ± 0.010 μmol mg−1 | PDT/PTT | 59 |
VONc | Two-step encapsulation | 0.256 ± 0.030 μmol mg−1 | ||||
2019 | COF-LZU1 | BODIPY-2H | BDF | 0.1545 ± 0.0220 mmol g−1 | PDT | 50 |
BODIPY-2I | 0.1360 ± 0.0312 mmol g−1 | |||||
2019 | TAPB-DMTP-COF | DOX | One-pot encapsulation | 32.1 wt% | Chemotherapy | 42 |
2019 | PI-CTF | Sorafenib | Two-step encapsulation | 83 wt% | Anti-cancer in vitro | 60 |
2019 | TP-Por-COF | CAD/IR783 | Two-step encapsulation | — | CT/PTT | 61 |
2020 | DT-COF | Carboplatin | Two-step encapsulation | 31.32 wt% | — | 62 |
2020 | TPB-DHTP-COF | 5-FU | Two-step encapsulation | 44 μg mg−1 | Chemotherapy | 63 |
COF-HQ | PSM | 73 μg mg−1 | ||||
2020 | SS-COF | DOX | Two-step encapsulation | 36 wt% | Chemotherapy | 64 |
F68@SS-COF | 21 wt% | |||||
2020 | TPA-TA-BD-COF | DOX | Two-step encapsulation | 35 wt% | Anti-cancer in vitro | 65 |
2020 | HY/SS COF | DOX | Two-step encapsulation | 18 wt% | — | 66 |
2020 | COFTTA-DHTA | Pirfenidone | Two-step encapsulation | 32 ± 4 wt% | PDT | 67 |
2021 | Tph-Dha-COF | Fe | Covalent binding | 1.6 mg mg−1 | Chemotherapy/CDT | 52 |
DOX | Two-step encapsulation | |||||
2021 | TA-COF | Ce 6 | Two-step | 14.3 wt% | Chemotherapy/PDT | 68 |
TPZ | encapsulation | 17.8 wt% | ||||
2021 | Tph-Dha-COF | Gambogic acid | Two-step encapsulation | 0.004 μg μg−1 | PTT | 69 |
2021 | Tf-TAPB-COF | FcCHO | Two-step encapsulation | 0.170 ± 0.018 μmol mg−1 | CDT | 51 |
0.201 ± 0.032 μmol mg−1 | ||||||
RSL3 | ||||||
2021 | TCOF | DOX | Two-step encapsulation | ∼28.57% | Chemotherapy | 70 |
2021 | RT-COF-1 | Lonidamine | Two-step encapsulation | 0.66 ± 0.05 mmol g−1 | Chemotherapy | 41 |
2021 | JUC-556 (E/Z) | Cytarabine | Two-step encapsulation | — | — | 71 |
2022 | DMTP-TAPB-COF | Heteropoly blue | One-pot encapsulation | — | PTT | 72 |
2022 | JUC-580 | Cisplatin | Two-step encapsulation | 20 wt% | Cell fluorescence imaging | 73 |
JUC-581 | ||||||
2022 | COF-366 | Plumbagin | Two-step encapsulation | 16.3 wt% | Anti-cancer in vitro | 74 |
2022 | TAPB-DMTA-COF | Curcumin (CUR) | One-pot encapsulation | 27.68 wt% | Wound dressing | 43 |
2022 | TRIPTA-COF | Cis | Two-step encapsulation | 31.19% | Chemotherapy | 39 |
2022 | DF-TAPB-COF | 5-FU | Two-step encapsulation | up to 60 wt% | — | 75 |
DF-TATB-COF | Captopril | |||||
2022 | TPE-ss COF | Matrine | Two-step encapsulation | 15.61 ± 0.23 wt% | Myocardial ischemia/reperfusion injury | 76 |
2022 | AQ4N@THPPTK-PEG | AQ4N | Two-step encapsulation | 26.35 wt% | CT/PDT | 77 |
2022 | TAPP-ANT-COF | Cypate | Two-step encapsulation | 11.2% | PDT/PTT | 78 |
2022 | PER@PDA-COF-1 | Mitoxantrone | Two-step encapsulation | ∼63% | CT | 79 |
2022 | DiSe-Por | DOX | Two-step encapsulation | 35.12% | CT/PTT/CDT | 80 |
2022 | COF-909 | Cu | Coordinated bonding | 10.76% | CDT/PDT/Immunotherapy | 54 |
Fe | 10.05% | |||||
Ni | 9.17% | |||||
2022 | TAPB-DMTP-COF | DOX | Two-step encapsulation | 92 μg mg−1 | Tumor therapy | 81 |
Camptothecin | 222 μg mg−1 | |||||
2022 | APTES-COF-1 | Rapamycin | Two-step encapsulation | — | Attenuate retinal ganglion cells death | 82 |
2022 | Hollow PDA-TMD-COF | Apatinib | Two-step/hollow encapsulation | ∼40.5 wt% | MWTT/CT | 83 |
2022 | TUS-84 | IBU | Two-step encapsulation | 11.05 wt% | — | 38 |
2022 | SP-COFs | POCl3 | Covalent binding | — | Inhibition of amyloid-β (Aβ) fibrillation | 84 |
2023 | TUS-64 | Captopril/IBU/isoniazid/5-FU/brimonidine | Two-step encapsulation | 4.09–17.37 wt% | — | 85 |
2023 | DSPP-COF | 5-FU | Two-step encapsulation | ∼0.86 μmol mg−1 | Chemotherapy/PDT | 86 |
2023 | DMTP-TPB-COF | ICG/AQ4N | Two-step encapsulation | 53.1% for ICG | PTT/PDT/chemotherapy | 40 |
2023 | Hollow HAPTP-TFPB/TPA/TFPA 2DCOFs | IBU | Two-step encapsulation | 12–20 wt% | — | 45 |
2023 | TAPB-DMTP-COF | Rose Bengal (RB) | Two-step encapsulation | 20.6% | SDT | 87 |
2023 | TAPB-DVA-COF | DOX/RB/CUR/ICG/Ce6 | Hollow encapsulation | 41.7–67.8% | SDT | 47 |
2015 | COF-DhaTab | Trypsin | Two-step encapsulation | 15.5 μmol g−1 | — | 88 |
2018 | NKCOF-1 | Lysozyme/tripeptide/lysine | Covalent binding | 22 μmol g−1 for lysozyme | Chiral separation | 89 |
2018 | TPB-DMTP-COF | Amano lipase PS | Two-step encapsulation | 0.95 mg mg−1 | — | 90 |
2020 | TPMM COF | α-amylase | Two-step encapsulation | 550 mg g−1 | Biocatalyst | 91 |
2020 | Tph-Dha-COF | Survivin antisense strand | Two-step encapsulation | 0.572 nmol mg−1 | Tumor imaging/PDT/prognostic evaluation | 92 |
2020 | TAPB-BTCA-COF | Ovalbumin | Two-step encapsulation | — | PTT/immunotherapy | 93 |
2020 | COF-1 | GOx/iunsulin | Two-step encapsulation | ∼45% | Anti-type 1 diabetes mellitus | 94 |
COF-5 | ||||||
2021 | TTA-DFP-nCOF | Insulin | Two-step encapsulation | ∼65 wt% | Anti-type 1 diabetes mellitus | 95 |
2021 | DHA-TAPP-COF | GOx | Two-step encapsulation | 46.1 μg mg−1 | PDT/starvation therapy | 96 |
CAT | 392.58 μg mg−1 | |||||
2021 | TPB-DMTP-COF | Survivin mRNA | Two-step encapsulation | 0.106 nmol mg−1 | Biomarkers fluorescence imaging | 97 |
MUC1 aptamers | 0.120 nmol mg−1 | |||||
2021 | Tph-Dha-COF | DNA | Covalent binding | 0.695 nmol mg−1 | Cancer cell imaging | 98 |
2021 | CTF-PEG-PEI | pDNA | Two-step encapsulation | — | Gene transfection | 99 |
2022 | COF-LZU1 | Trypsin/HRP/GOx/BSA | One-pot encapsulation | 5.0 mg g−1 for trypsin | — | 44 |
2022 | [OH]x%-TD-COFs | Cytochrome c (Cyt c) | Covalent binding | — | — | 100 |
2022 | BPTA-TPB-COF | Peptide/HRP | Covalent binding | — | Protein detection | 101 |
2022 | ETTA-TPAL-COF | Glucose oxidase GOD | Two-step encapsulation | 30 μg mg−1 | Glucose detection | 102 |
2022 | Alkynyl-COF | DNA | Covalent binding | 16 μg mg−1 | Exosomes detection | 103 |
2023 | Alkynyl-COF | DNA/HRP | Covalent binding | — | Exosomes analysis | 104 |
2023 | TAPB-DMTA-COF | Peptide T5/SOD/CAT | Covalent binding | 16.29 wt% | Modulates oxidative stress in Alzheimer's disease | 105 |
Two-step encapsulation | ||||||
2023 | Tp-Bd-COF | Cyt c/GOx/HRP | Hollow encapsulation | 57–127 mg g−1 | Glucose detection | 106 |
2023 | TAPB-DMTA COF | GOx/Cyt c | Two-step encapsulation | — | Biocatalyst | 107 |
2022 | TPB-DHTA-COF | Exosome | Two-step encapsulation | — | Diabetic fester wound healing | 108 |
For example, a three-dimensional (3D) COF (TUS-84) with scu-c topology was synthesized via the [8+4] imine condensation reaction in 2022.38 This is a typical example of a COF synthesized through the traditional solvothermal method, where two monomers (DPTB-Me and TAPP), catalyst (6 M aqueous acetic acid) and solvent (mesitylene and 1,4-dioxane, 5:5, v/v) are mixed at 120 °C for 3 days. As shown in Fig. 2(A), the PXRD and HRTEM characterization results of TUS-84 indicated it had an ordered microporous structure with high crystallinity, which showed great potential for drug delivery. Subsequently, in the second step of drug loading, the TUS-84 powder was immersed in a hexane solution of ibuprofen, and then stirred at room temperature for 4 h. Both UV-Vis spectroscopy and thermogravimetric analysis (TGA) were conducted to measure the loading capacity of ibuprofen in Ibu@COF, and the results from these two methods corresponded well, revealing a value of 11 wt%. Similarly, Sabuj Kanti Das and co-workers employed a novel biocompatible COF (TRIPTA-COF) to serve as a great nanocarrier for cisplatin (Fig. 2(B)).39 Briefly, the monomers were added to a dry Pyrex tube, followed by the addition of the solvent and catalyst. Subsequently, the homogeneously dispersed reaction mixture was degassed by three freeze–pump–thaw cycles, flame-sealed, and then allowed it to react for 4 days. Finally, cisplatin was impregnated in the unoccupied pores of the COF by overnight stirring using DMSO as the solvent. The quantitative drug loading assay was performed by TGA and ICP-OES techniques, resulting in a high loading efficiency of 31.19%.
Fig. 2 The two-step encapsulation procedure via the solvothermal method. (A) I. Strategy for the constructing of 3D COFs with scu or scu-c topology. II. PXRD patterns and HRTEM images of TUS-84. Reproduced from ref. 38 with permission from ACS Applied Materials & Interfaces, Copyright 2022. (B) Synthesis of cisplatin-loaded TRIPTA-CISPLATIN. Reproduced from ref. 39 with permission from Nanoscale Advances, Copyright 2022. |
Unlike traditional solvothermal COF synthesis, room temperature synthesis approaches do not require any rigorous reaction conditions, such as high temperature, long reaction time and inert atmosphere (absence of water and oxygen). Alternatively, mild reaction conditions are employed to construct cargo@COF systems. In 2023, Han et al. synthesized a dual-drug delivery system to immobilize both the photosensitizer indocyanine green (ICG) and the hypoxia-activated prodrug AQ4N.40 In a typical synthesis procedure, a transparent homogeneous solution of monomers and catalyst was stirred for 12 h at room temperature. Subsequently, the residue was collected by centrifugation, followed by washing with anhydrous ethanol three times. The pre-synthesized TPB-DMTP-COF was used to entrap ICG and AQ4N stepwise through stirring for 12 h to obtain COF@ICG and COF@ICG/AQ4N, respectively. As shown in Fig. 3(A), the TEM images verified that all products were dispersed uniformly.
Fig. 3 The two-step encapsulation procedure with room temperature synthesis approach. (A) I. Synthesis process of HA-COF@ICG/AQ4N. II. TEM images of COF, COF@ICG, and HA-COF@ICG/AQ4N nanoparticles. Reproduced from ref. 40 with permission from Colloids and Surface B: biointerfaces, Copyright 2023. (B) I. Synthesis process of imine-linked COFs. II. SEM images of COFs and equipment for the gram-scale synthesis of RT-COF-1. III. Synthesis process of LND@RT-COF-1. Reproduced from ref. 41 with permission from Advanced Therapeutics, Copyright 2021. |
In almost all previous studies, COFs were synthesized on the milligram-scale; however, Dong's group proposed a novel strategy that expanded the synthesis procedure to the gram-scale, while maintaining the particle size and spherical morphology.41 The photocatalytic cascade reactions played a dominant role in the newly propounded strategy, which were performed to prepare four previously reported imine-linked COFs after sequential reactions under ambient conditions at room temperature (Fig. 3(B)). Subsequently, RT-COF-1 with an average size of 167.9 nm was selected as a model nanocarrier for the delivery of the anticancer drug lonidamine (LND). As shown in Fig. 3(B)-III, the LND@ RT-COF-1 composite was simply obtained by immersing RT-COF-1 in an ethanol solution of LND for 24 h.
Fig. 4 The one-pot encapsulation method. (A) Illustration of the synthesis and application of DOX@COF. Reproduced from ref. 42 with permission from Chemistry, Copyright 2019. (B) A schematic diagram of the preparation of CUR@COF. Reproduced from ref. 43 with permission from ACS Applied Materials & Interfaces, Copyright 2022. (C) The synthesis process and formation mechanism of the trypsin@COF-LZU1 capsule. Reproduced from ref. 44 with permission from ACS Applied Materials & Interfaces, Copyright 2022. |
Recently, Zou et al. fabricated a CUR@COF nanosystem via a facile “one pot” method for the delivery of curcumin (CUR).43 The CUR loading capacity in the COF was calculated to be in the range of 15.05% to 27.68% with an increase in the content of CUR from 2 to 8 mg, while the encapsulation efficiency decreased from 78.26% to 49.63%. Furthermore, the standard CCK-8 assay was conducted to evaluate the in vitro cytotoxicity of COF or CUR@COF, and the results showed that 91% and 95% cells survived at the concentration of 200 μg mL−1, indicating the low cytotoxicity and good biocompatibility of COF and CUR@COF (Fig. 4(B)).
The “one-pot” method was used for the delivery of not only small molecules but also for the immobilization of biomacromolecules, such as enzymes. Recently, a hollow COF-based capsule for the template-free in situ encapsulation of enzymes was prepared by Chao and co-workers.44 As shown in Fig. 4(C), the enzyme (trypsin) and raw materials in process for the synthesis of COF-LZU1 (benzene-1,3,5-tricarbaldehyde and p-phenylenediamine) were mixed in 1,4-dioxane at room temperature and left to react for 3 days without stirring. Finally, the trypsin@COF-LZU1 composites were successfully prepared via the single-step and template-free process. With an increase in the content of trypsin from 0 to 6.0 mg, the crystallinity of COF-LZU1 showed no significant change given that the X-ray diffraction peaks of the products remained the same. More importantly, this single-step strategy can also be extended to other biomolecules (HRP, GOx and BSA), making it a high potential method for the encapsulation of biomacromolecules.
Recently, Du and co-workers proposed a self-templating method for the synthesis of a hollow 2D COF using a newly synthesized building block (2,3,6,7,10,11-hexakis(4-aminophenyl) triphenylene, HAPTP) under solvothermal conditions.45 Impressively, the newly synthesized COFs were homogeneous solid spheres initially; however, a transformation occurred based on a time-dependent model, gradually turning into uniform hollow spheres accompanied by the Ostwald ripening process, as shown in Fig. 5(A). Subsequently, IBU was loaded successfully, and the loading capacity was determined by TGA to be as high as 20 wt%, benefitting from the high crystallinity quality and large surface area of the host COFs.
Fig. 5 The hollow encapsulation procedure. (A) I. Process for the synthesis of HAPTP-TPA 2DCOFs. Reproduced from ref. 45 with permission from Chemistry of Materials, Copyright 2023. (B) Synthetic route for biomacromolecule@COF capsules. Reproduced from ref. 46 with permission from Journal of the American Chemical Society, Copyright 2020. (C) Synthesis process of the COF and hollow COF. Reproduced from ref. 47 with permission from Angewandte Chemie-International Edition, Copyright 2023. |
Li et al. used metal–organic frameworks (MOFs) as sacrificial templates to construct COF capsules for the encapsulation of an enzyme (Fig. 5(B)).46 Firstly, the catalase (CAT) molecules were in situ encapsulated in digestible MOFs to form biomacromolecules@MOF, which could protect enzymes from losing their bioactivities. Secondly, the COF coating was grown on the surface to obtain biomacromolecule@MOF@COF core–shell structures. Finally, the MOF core was etched and the as-obtained hollow COF was used to immobilize the enzyme successfully. Consequently, compared with free CAT or the CAT@MOF complex, the CAT@COF capsules exhibited an excellent protection effect on the enzyme activity under various perturbation conditions including acid, proteases, acetone and high temperature.
Liu and coworkers proposed a fascinating strategy for the controllable synthesis of a series of hollow COFs via the oxidation of the imine bond via hydroxyl radicals (˙OH) generated from the Fenton reaction.47 The whole procedure included two steps (Fig. 5(C)), where firstly, a solvent evaporation strategy was proposed for the synthesis of high crystallinity COFs, by which the solvent (ACN) and catalysis (acetic acid) were evaporated when exposed in an uncovered vessel at mild temperature (room temperature to 60 °C) with stirring. Secondly, the synthesized COFs were converted into hollow COFs by breaking the imine bond under the combined action of ferric chloride (FeCl3), hydrogen peroxide (H2O2) and acetic acid. Whether in step one or two, several COFs or HCOFs with diverse morphologies were successfully achieved, demonstrating the universality of this strategies. Finally, the synthesized TAPB-DVA-HCOF functioned as a nanocarrier to load five small molecule drugs. The DLC was calculated to be in the range of 41.7% to 67.8% and EE calculated to be in the range of 31.5% to 75.1% for the five drugs, which were much higher than previous reports. Finally, Ce6@HCOF showed an excellent enhanced sonodynamic cancer therapy ability and anti-tumor efficiency both in vitro and in vivo.
The active sites of COFs can be connected with specific guest molecules through covalent bonds. For example, Shouvik Mitra et al. covalently connected folic acid (FA) to a COF (TpASH-FA) through three sequential post-synthetic modifications (Fig. 6(A)).48 Firstly, the phenolic hydroxyl groups of the original COF (TpASH) were used to conjugate glycidol (Glc). Secondly, the surface hydroxy groups of Glc were used for conjugation of 3-aminopropyltriethoxysilane (APTES) to produce amine (–NH2)-functionalized COFs. Thirdly, the free amino groups of the other end of the APTES molecules reacted with the carboxyl groups of FA through the EDC/NHS coupling reaction, resulting in the formation of the folate-conjugated targeted COF-based delivery system (TpASH-FA). This was the first case of COF-based targeted drug delivery using sequential post-synthetic modifications.
Fig. 6 Post-synthetic modification (PSM). (A) I. A schematic diagram of the chemical structure of COF nanosheets. II. Sequential PSM processes of TpASH to TpASH-FA and their NMR spectra (III), as well as SEM and TEM images (IV). Reproduced from ref. 48 with permission from Journal of the American Chemical Society, Copyright 2017. (B) I. Two reversible redox states of a bipyridine-like structure. II. Sequential PSM processes of Py-BPY-COF to Py-Bpy2+-COF or Py-BPy+˙-COF. Reproduced from ref. 49 with permission from Journal of the American Chemical Society, Copyright 2019. |
Besides the active sites of the linkers, linkers with particular structures can also be target locations for PSM. Guo's team fabricated a 2,2′-bipyridine-based COF using sequential reactions in situ.49 The obtained Py-BPy-COF contained a 2,2′-bipyridine-like structure whose dicationic derivatives could undergo two-step reversible reductions to form a radical cation and a neutral species (Fig. 6(B)-I), leading to the formation of two cationic types of COFs, Py-Bpy2+-COF and Py-BPy+˙-COF (Fig. 6(B)-II).
The bonding defect functionalization (BDF) was proposed by Dong's group in 2019, which is an alternative strategy to covalently connect the guest with the host.50 In practical terms, the unbonded functional groups (bonding defects) at the end of the COF matrix are supposed to be ideal active sites for grafting small organic molecules (guest). For example, Dong's group prepared two boron-dipyrromethene (BODIPY)-decorated nanoscale COFs (NCOFs) through this BDF approach, i.e., a Schiff-base condensation between the free end aldehyde groups (bonding defects in COFs) of NCOFs and the amino moieties of the organic photosensitizer BODIPY (Fig. 7(A)).50 Similarly, through the BDF method, Dong's group also fabricated a COF-Fc nanosystem (Fig. 7(B)) through the Schiff-base condensation between the free end amino groups (bonding defects in COFs) of COFs and the aldehyde moieties of the Fenton-like reaction catalyst ferrocenecarboxaldehyde (FcCHO).51
Fig. 7 Bonding defect functionalization (BDF). (A) Design of BODIPY-loaded nanoscale COFs and their photographs. Reproduced from ref. 50 with permission from iScience, Copyright 2019. (B) Material design and synthesis process of RSL3@COF-Fc. Reproduced from ref. 51 with permission from Small, Copyright 2021. |
Fig. 8 Coordination bonding strategy. (A) Synthetic processes of DOX@COF(Fe) and its application. Reproduced from ref. 52 with permission from Chemical Communications, Copyright 2021. (B) Synthetic processes of CFAP and its application. Reproduced from ref. 53 with permission from ACS Applied Materials & Interfaces, Copyright 2022. (C) I. Illustration of the construction of bulk species (L-3N-M) or porous scaffolds (COF-909-M). II. Schematic diagram of multienzyme-mimicking COFs. Reproduced from ref. 54 with permission from Advanced Materials, Copyright 2022. |
In 2020, Liu et al. fabricated a TAPB-DMTP-COF using two types of monomers, 1,3,5-tris(4-aminophenyl)benzene (TAPB) and 2,5-dimethoxyterephthaldehyde (DMTP).53 Subsequently, to endow the platform with the ability to further catalyze and degrade the overexpressed H2O2 in the acidic tumor microenvironment (TME) through Fenton-like reactions, ferric chloride was introduced, by which Fe3+ could be coordinated with the amino groups of the COFs to form metallized COFs (Fig. 8(B)). The pre-synthesized CFAP nanocomposite showed efficacious synergistic photo-, chemodynamic-, and immunotherapy effects towards both the primary tumor and tumor metastasis.
In addition, COF-909-Cu with a bipyridine-like structure monomer is another typical example of metal coordination, which exhibited excellent CDT efficacy, and also efficiently triggered pyroptosis.54 The adjacent tri-pyridines of 4,4′,4′′-(1,4-phenylene)bis(([2,2′:6′,2′′-terpyridine]-5,5′′-dicarbaldehyde)) were ideal candidates for synthesizing metalated COFs. The tri-pyridine structure acted as a paw, which anchored the metal ions (e.g., Cu2+, Fe3+, and Ni2+) tightly, as illustrated in Fig. 8(C).
In 2015, Fang and co-workers innovatively proposed that COFs have potential for drug delivery for the first time (Fig. 9(A)).55 They fabricated two PI-based COFs, which were denoted PI-COF-4 and PI-COF-5, using a linear dianhydride and two different tetrahedral amines through an imidization reaction, exhibiting high crystallinity and considerable BET surface areas. The selected PI-COFs had pore sizes in the range of 11 Å to 15 Å, which provided sufficient room to entrap ibuprofen (IBU) molecules, which have a size of approximately 5 Å × 15 Å. Furthermore, the loading capacity of IBU was evaluated to be 24 wt% for PI-COF-4 and 20 wt% for PI-COF-5 based on the TGA analysis.
Fig. 9 Small molecule@COF. Synthesis process of (A) PI-COF-4 and PI-COF-5. Reproduced from ref. 55 with permission from the Journal of the American Chemical Society, Copyright 2015. (B) PI-3-COF and PI-2-COF. Reproduced from ref. 56 with permission from Chemical Communications, Copyright 2016. (C) DOX-loaded PEG-CCM@APTES-COF-1. Reproduced from ref. 58 with permission from Nature Communications, Copyright 2018. (D) Synthesis process of TD-COFs and SP-COFs. Reproduced from ref. 84 with permission from Chemical Science, Copyright 2022. |
In the next year, compared with PI-COF-4 or PI-COF-5 on the micro-scale, nano-sized COFs were successfully fabricated, realizing tremendous progress for biomedical application. Bai et al. used 1,3,5-triformylbenzene and two types of amine compounds to synthesize two other polyimine-based COFs, namely, PI-3-COF and PI-2-COF (Fig. 9(B)).56 Subsequently, by following one gold standard that the molecular size of the guest should fit well with the pore width of host, three different drug molecules (5-FU, captopril and IBU) were selected, resulting in a high DLC, which reached up to 30 wt%. Unlike the micro-scale PI-COF-4 and PI-COF-5 synthesized by Fang's group, the average diameter of PI-3-COF and PI-2-COF was about 50 nm, providing a greater possibility for applications in vivo. Moreover, the phenomenon that the drug-loaded COFs can be endocytosed in cancer cells was observed and confirmed by confocal laser scanning microscopy, providing direct evidence for nano COFs serving as drug carriers.
Subsequently, Guiyang Zhang et al. developed an amine-functionalized COF-1 (APTES-COF-1) to encapsulate DOX, which is a classic anticancer drug,58 followed by decorating with fluorescent composite polyethylene-glycol-modified monofunctional curcumin derivatives (PEG-CCM), and then obtained a water-dispersible polymer-COF-DOX nanocomplex (Fig. 9(C)). Notably, this was the first time that COF-based nanocomposites were investigated for in vivo drug delivery and antitumor efficacy in nude mice, which can facilitate innovations and applications for COF-based carriers for drug delivery.
Recently, Yao and co-workers used this type of PSM strategy to functionalize TD-COFs with sodium phosphate groups in their channels to obtain phosphorylated COFs (SP-COFs), as shown in Fig. 9(D).84 Interestingly, the hydroxyl unit ligands of TD-COFs were ideal phosphorylation sites, and it only took two simple steps to obtain the final product SP-COFs. After exploring the mechanism of action between SP-COFs and the Aβ42 peptide (a typical cytotoxin to Alzheimer's disease) through molecular dynamics simulation, the single pore of SP-COFs was proven to have potential for the recognition and targeting of specific amino acid sites of the Aβ42 sequence. Besides, the following cellular level experiments also demonstrated that SP-COFs are new nano-inhibitors for the inhibition of Aβ fibrillation for the prevention and treatment of Alzheimer's disease.
In addition to drug molecules, COFs can also serve as carriers for photosensitizers (PSs) and photothermic agents (PTAs), where the former are employed photodynamic therapy (PDT), while the latter for photothermal therapy (PTT). Guan et al. prepared a nano COF-based system (VONc@COF-Por) through a facile synthetic approach under ambient conditions.59 In this platform, porphyrinic PS (Por) was covalently grafted through the bonding defect functionalization strategy and the naphthalocyanine PTA (VONc) was noncovalently loaded via the guest encapsulation approach to obtain a multifunctional nanosystem for PDT/PTT combination therapy, as shown in Fig. 10(A). Quantitatively, the loading capacity of Por and VONc was determined by the UV-Vis external standard method and calculated to be 0.091 ± 0.010 and 0.256 ± 0.030 μmol mg−1, respectively.
Fig. 10 Small molecule@COF. (A) Mind mapping for designing nanotherapeutic systems and synthesis process of VONc@COF-Por. Reproduced from ref. 59 with permission from ACS Nano, Copyright 2019. (B) Schematic illustration of synthesis of COF@IR783@CAD and its application. Reproduced from ref. 61 with permission from ACS Applied Materials & Interfaces, Copyright 2019. (C) Fabrication process and photoimmunotherapy performance of ICG@COF-1@PDA nanosheets. Reproduced from ref. 109 with permission from Advanced Functional Materials, Copyright 2019. (D). The synthetic process of RCMP and its application in SDT. Reproduced from ref. 87 with permission from Advanced Science, Copyright 2023. |
Various COF-based delivery systems for PSs and PTAs have successively emerged. Wang et al. developed the COF@IR783@CAD dual-delivery system for the combination of chemotherapy and PTT (Fig. 10(B)).61 More importantly, this complex showed an excellent photo-absorption property and good light-to-heat conversion efficiency, suggesting that COF@IR783 can be used for photoacoustic (PA) imaging in vivo. Besides, Gan et al. constructed COF-1 using 1,4-benzendiborinic acid as a monomer for the delivery of ICG in 2019 (Fig. 10(C)).109 Besides, Ge et al. obtained a TA-COF, which was synthesized using 1,3,5-triformyl-2,4,6-trihydroxybenzene (TP) and 4,4-azodiaminobenzene (AD) for the delivery of Ce6 in 2021.68
Recently, Zhang and co-workers designed a COF nanobowl and engineered it as an activatable nanosensitizer for the delivery of the small-molecule sonosensitizer Rose Bengal (RB) to realize activatable and ferroptosis-boosted sonodynamic therapy (SDT).87 As shown in Fig. 10(D), COF nanobowls with high crystallinity and uniform morphologies were synthesized using a hard-template method. Benefiting from the MnOx shell with glutathione (GSH)-responsive degradation behavior, the nanoplatform could switch from the “off” to “on” state to exert therapeutic efficacy at the tumor site under US irradiation.
Fig. 11 Protein@COF. (A) Preparation of the COF@GOx@CAT nanopocket for cancer therapy. Reproduced from ref. 96 with permission from Chemical Communications, Copyright 2021. (B) The fabrication and mechanism of phototherapy and immunotherapy by CIO. Reproduced from ref. 93 with permission from Journal of Materials Chemistry B, Copyright 2020. (C) Covalent immobilization of Cyt c onto COFs. Reproduced from ref. 100 with permission from Angewandte Chemie-International Edition, Copyright 2022. (D) Synthesis of FITC-PEG-COF@Ins-GOx composites (I and II) and the possible mechanism of glucose and pH dual-responsive insulin delivery (III). Reproduced from ref. 94 with permission from ACS Applied Materials & Interfaces, Copyright 2019. |
In addition, ovalbumin (OVA) is a common antigen used for antigen-induced immunotherapy. Pang's group developed a COF-based antigen delivery system (named COF@ICG@OVA, CIO) for synergistic cancer phototherapy and immunotherapy (Fig. 11(B)).93 The particle size of the composite material was measured to be ∼100 nm, and the loading efficiency towards OVA was around 40.44%. More importantly, the CIO nanoparticle could not only ablate primary tumors through PDT/PTT, but also induce antitumor immune responses combined with checkpoint blockade therapy.
Recently, Xing et al. covalently immobilized cytochrome c (Cyt c) in [OH]x%-TD-COFs.100 These COFs were synthesized using one triamine 1,3,5-tris(4-aminophenyl) benzene (TAPB) and two dialdehydes including 2,5-dihydroxyterephthalaldehyde (DHTA) and 2,5-dimethoxyterephthalaldehyde (DMTA) with different proportions (Fig. 11(C)), where x% represents the molar percentage of DHTA in the dialdehyde mixture. Therefore, the obtained COFs had different contents of anchoring sites (phenolic groups) for further introducing epoxy units on the channel walls of the COFs through a Williamson ether reaction. The third step was to anchor Cyt c in the COFs by the covalent linkage between nucleophilic groups from the enzyme guest (mainly the amino moiety) and epoxy groups from the COF host. Consequently, the covalent interactions between COFs and Cy c could induce secondary structural changes in the latter, resulting in a more accessible heme center, and thus enhancing the catalytic activity, which could reach 600% activity compared with the free enzyme.
Zhang et al. firstly prepared two borate-based COFs (COF-1 and COF-5) for the encapsulation of insulin (Ins) and glucose oxidase (GOx) via Brønsted and Lewis type (N: → B) complexation, and subsequently decorated them with polyethylene glycolated isothiocyanate (FITC-PEG) to finally obtain FITC-PEG-COF@Ins-GOx (Fig. 11(D)).94 The obtained composite was suggested to achieve the glucose- and pH dual-responsive delivery of COF-based insulin. Due to the presence of fluorescein FITC, the bio-distribution of the composite could be seen by tracking the intrinsic fluorescence of FITC in vivo. Besides, this composite showed excellent anti-diabetic effects on type I diabetic mice and could maintain the blood glucose level in the normal range.
In addition to intravenous delivery, COFs could also be used for the oral delivery of insulin. Farah Benyettou and co-workers prepared layered 2D COF nanosheets with insulin loaded between the nanosheet layers (TTA-DFP-nCOF/insulin) with a high loading capacity (∼65 wt%).95 It was noteworthy that TTA-DFP-nCOF could protect the encapsulated insulin to maintain its activity under harsh conditions mimicking the stomach environment (pH = 2.0); meanwhile, the sustainable release of insulin could be successfully implemented under hyperglycemic conditions.
Fig. 12 Nucleic acid@COF. (A) Synthesis of the COF–survivin system. Reproduced from ref. 92 with permission from Chemical Science, Copyright 2020. (B) Construction of CTF-PEG-PEI for efficient intracellular gene delivery. Reproduced from ref. 99 with permission from ACS Applied Nano Materials, Copyright 2021. (C) Construction of CLZU nanoparticles and their application for anti-tumor gene therapy. Reproduced from ref. 110 with permission from Science Chins Chemistry, Copyright 2021. (D) Construction of DNA-COFs for the detection of exosomes. Reproduced from ref. 103 with permission from Analytical Chemistry, Copyright 2022. |
Cao et al. reported a strategy to construct triazine-based COF nanosheets (CTF-PEG-PEI) via exfoliation and surface modification for the delivery of DNA (Fig. 12(B)).99 This nonviral vector had a unique brush-like hierarchical structure, making it strongly condense nucleic acids and efficiently release the nucleic acid cargo after cellular uptake. To further study the in vivo gene delivery ability of COFs, Hao et al. designed a series of cationic porous COF-based nanoparticles with a good gene transfection effect and biocompatibility (Fig. 12(C)).110 After successfully constructing the vehicles, a therapeutic gene for tumor suppression, shVEGF, was chosen for in vivo delivery. Finally, this platform could significantly inhibit tumor growth, indicating its excellent potential for gene therapy.
Yang's group firstly proposed a strategy to functionalize COFs with DNA by covalent interaction.103 As shown in Fig. 12(D), the azide-modified DNA could be efficiently connected to the pre-prepared alkynyl-rich COFs through Cu(I)-catalyzed azide/alkyne cycloaddition (CuAAC) reaction. Then, a complementary 5-carboxyfluorescein-labeled DNA (FAM-DNA) strand was used to hybridize with the single-strand DNA on the COF surface to obtain DNA@COF composites. In this design, the COF probes were further used to construct an effective electrochemical biosensor for the detection of exosomes.
Fig. 13 Exosome@COF. I. Formation of TPB-DHTA-COF, synthesis of PCOF, and preparation of engineered exosome (E-Exo). II. TEM image of the PCOF@E-Exo nanoagent. III. Preparation of the PCOF@E-Exo integrated nanoagent and its theranostic mechanism in infected diabetic fester wounds. IV. Antibacterial activity analysis of the PCOF@E-Exo nanoagent against E. coli and S. aureus. Reproduced from ref. 108 with permission from Small, Copyright 2022. |
Year | Composite | Hybrid type | Cargo | Applications | Ref. |
---|---|---|---|---|---|
2019 | ICG@COF-1@PDA | PDA@COF | ICG | Immunotherapy/PTT | 109 |
2020 | HRP-pSC4-AuNPs@COFs | Au NP@COF | HRP | Exosomes detection | 112 |
2023 | ICPA | Pt NP@COF | ICG | Self-strengthening photocatalytic therapy | 113 |
2019 | Co-MOF@TPN-COF | MOF@COF | Aptamer | Ampicillin detection | 114 |
2021 | Cu-MOF@CuPc-TA-COF | MOF@COF | pDNA | HIV-1 DNA detection | 115 |
2021 | Cu-MOF@TpBD-COF | MOF@COF | Aptamer | PDGF-BB detection | 116 |
2022 | MOF@COF-Apatinib | MOF@COF | Apatinib | MWDT/anti-angiogenesis therapy | 117 |
2022 | CA4V/ZIF-90@TzCOF@Apt | MOF@COF | CA4V | Anti-vascular therapy | 118 |
2022 | C60@TAPT-DHTA-COF | C60@COF | Aptamer | Tobramycin detection | 119 |
2022 | O2-FeCOF@CaCO3@FA | CaCO3@COF | FA | PDT/Ca2+ overload treatment | 120 |
2021 | Au/COF/MnO2 | MnO2@COF@Au NP | Antibody | HCG detection | 121 |
2021 | Fe3O4@COF-BSA–FA | Fe3O4@COF | DOX | PTT/CT | 122 |
2021 | Fe3O4@COF@BSA | Fe3O4@COF | BSA | Chiral recognition | 123 |
2022 | Fe3O4@SiO2@COF | Fe3O4@SiO2@COF | Hairpin DNA | Glioma detection | 124 |
2022 | Fe3O4@COF-Au NP | Fe3O4@COF@Au NP | ssDNA | ATP detection | 125 |
2023 | GOx–Fe3O4@COF | Fe3O4@COF | GOx | Removal of mycotoxins | 126 |
2023 | MnO2-Poly(I:C)@COF | MnO2@COF | Poly(I:C) | Immuno-sonodynamic therapy | 127 |
2022 | MGPPCLP | SiO2@COF | Gemcitabine/glycoprotein | Programmed multi-drugs delivery | 128 |
Inhibitor/losartan | |||||
2022 | OC-COF | Cyclodextrin@COF | Ligustrazine (LIG) | Inhalation therapy against acute lung injury | 129 |
2023 | Por-COF-gel | Gel@COF | Acrylate | Sterilization/wound healing | 130 |
2023 | CuS@COFs-BSA–FA/DOX | CuS@COF | DOX/BSA | CT/PTT/CDT | 131 |
For example, Li et al. proposed the combination of a COF-based nanocapsule with an MOF (MOF@COF) for loading the hydrophobic inhibitor apatinib.117 The MOF as the inner layer consisted of Bi3+ and Mn2+, which was designed as a microwave sensitizer to generate cytotoxic 1O2 and heat for microwave dynamic/thermos synergistic therapy (MWDT/MWTT), as shown in Fig. 14(A). The COF as the outer layer was covalently coated on MOF for further augmenting these two treatment effects. Besides, Wang et al. proposed a bioorthogonal nanoreactor (CA4V/ZIF-90@TzCOF@Apt) with a ZIF-90 core and COF shell (Fig. 14(B)).118 Specifically, the ZIF-90 core was used for the delivery of CA4V, a prodrug of the vascular disrupting agent CA4, through a ‘one-pot’ in situ self-assembly process, involving CA4V, imidazole-2-carboxaldehyde (ICA) and Zn2+. Besides, the TzCOF shell was utilized for coating an aptamer polymer, which could target tumor cells.
Fig. 14 MOFs@COFs. (A) Synthesis of BMCAP nanocapsule and its application in MWDT/MWTT for anti-angiogenesis of colorectal cancer. Reproduced from ref. 117 with permission from Biomaterials, Copyright 2022. (B) Synthesis of CA4V/ZIF-90@COF@Apt nanoreactor and its significant cancer inhibition rates towards DU145 cells. Reproduced from ref. 118 with permission from Chemical Communications, Copyright 2022. |
Fig. 15 Metal nanoparticles@COFs. (A) I. Process for the fabrication of COF-based nanoprobes and II. mechanism of the electrochemical biosensor for the detection of exosomes. Reproduced from ref. 112 with permission from Biosensors and Bioelectronics, Copyright 2020. (B) Synthesis process of ICPA and its application in reversing tumor hypoxia and inducing tumor thermal damage and oxidative stress. Reproduced from ref. 113 with permission from Biomaterials, Copyright 2023. |
Recently, increasing efforts have been devoted to the combination of functional components, i.e., metal nanoparticle-doped COFs. Tang's group proposed the preparation of ultra-small Pt NPs deposited on a COF nanoplatform (termed CP) via an in situ chemical reduction method (Fig. 15(B)).113 Subsequently, a thiol-terminated AS1411 aptamer was decorated on it by forming a stable Pt–S bond and the photosensitizer ICG was loaded to obtain the final versatile nanosystem (ICPA).
Fig. 16 Metal oxides/sulfides@COFs. (A) Construction of MnO2-Poly(I:C)@COF NPs for enhanced SDT. Reproduced from ref. 127 with permission from Advanced Functional Materials, Copyright 2022. (B) Synthesis of Fe3O4@COF-BSA–FA system and its synergistic treatment. Reproduced from ref. 122 with permission from Microporous and Mesoporous Materials, Copyright 2021. (C) Preparation and therapeutic functions of CuS@COFs-BSA–FA/DOX. Reproduced from ref. 131 with permission from Chemical Engineering Journal, Copyright 2023. |
Zhao et al. integrated a magnetic COF with Fe3O4 shell and further functionalized it with bovine serum albumin (BSA) and folic acid (FA) to improve its dispersibility and selectivity.122 Finally, the prepared novel Fe3O4@COF-BSA–FA nanosystem worked as an effective DOX carrier with excellent photothermal performance, resulting in chemo/photothermal combined therapy (Fig. 16(B)). Similarly, Wang et al. established a multifunctional nanoplatform (CuS@COFs-BSA–FA/DOX) based on a hybrid CuS/COF inner layer and BSA/FA outer layer for synergistic chemo/photothermal/chemodynamic therapy (Fig. 16(C)).131
COF | Linkage | Stimuli type | Morphology and size | Cargo | Release behavior | Disease model | Ref. |
---|---|---|---|---|---|---|---|
TAPB-DMTP-COF | Imine | pH | Nanosphere (200–400 nm) | DOX | Almost 100% at pH = 5.0 in 24 h | H22 xenograft mouse model | 42 |
TAPB-DMTA-COF | Imine | pH | Multiantenna-like particle (160.5 nm) | Curcumin | 74.6% at pH = 5.0 in 100 h | Animal full-thickness wound defect model | 43 |
DT-COF | Imine | pH | Nanosphere (500–600 nm) | Carboplatin | Completely released for 1 h at pH = 5.0 and 12 h at pH = 7.4 | — | 62 |
Tf-TAPB COF | Imine | pH | Nanosphere (∼170 nm) | RSL3 | 61.1% ± 9.0% at pH = 5.0 in 96 h | HT-1080 xenograft mouse model | 51 |
TPB-DMTP-COF | Imine | pH | Radial sphere (144 nm) | Gambogic acid | 59.43% at pH = 5.0 in 24 h | 4T1 xenograft breast cancer model | 132 |
COF-366 | Imine | pH | Sphere (∼150 nm) | Plumbagin | 93% at pH = 5.5 in 72 h | LNCap cells | 74 |
Hollow COF | Imine | pH | Hollow nanosphere (∼30 nm shell thickness) | Apatinib | 76.3% at pH = 5.7 in 24 h | HepG2 xenograft mice | 83 |
TTA-DFP-nCOF | Imine | pH | Nanosheet (stacking of ∼18 layers) | Insulin | Almost 100% at pH = 7.4 with 5 mg mL−1 glucose after 7.5 h | Streptozotocin-induced type 1 diabetic rat model | 95 |
PI-CTF | Triazine | pH | Tubular structure (∼30 nm) | Sorafenib | 66% at pH = 5.3 in 48 h | LNCaP cells | 60 |
JUC-56-[HZ]x | Hydrazone | pH | — | Cytarabine | 74.56% at pH = 4.8 in 72 h | — | 71 |
F68@SS-COF | Imine and disulfide | pH/GSH | Nanosphere (140 ± 15 nm) | DOX | 90% at pH = 5.0 with 10 mM GSH in 24 h | HepG2 cells | 64 |
HY/SS-CONs | Imine and disulfide | pH/GSH | Nearly sphere (120 ± 20 nm) | DOX | About 90% at pH = 5.0 with 10 mM GSH in 72 h | HepG2 cells | 66 |
DSPP-COF | Imine and disulfide | pH/GSH | 59 nm | 5-Fu | ∼96.9% at 10 mM GSH in 24 h | MCF-7 xenograft nude mouse | 86 |
TPE-ss COF | Imine and disulfide | pH/GSH | Nanosheet (5 nm thickness) | Matrine | ∼80% at pH = 5.0 with 10 mM GSH in 80 h | Myocardial ischemia/reperfusion injury | 76 |
Dise-Por | Imine and diselenide | pH/GSH/photo | Sheet-like morphology | DOX | 89.6% at pH = 5.4 with 10 mM GSH and 808 nm laser in 96 h | PC-3 tumor xenografted mice | 80 |
TA-COF | Imine and azo | Hypoxia | Sphere-like (90 nm) | TPZ | ∼65% with 2 mM Na2S2O4 in 2 h | 4T1 tumor xenografted model | 68 |
PER@PDA-COF-1 | Imine | Protein | Nanosheet | Mitoxantrone | UV-Vis absorption increase and shift when adding 0.3 mM human serum albumin | CT-26 tumor xenografted model | 79 |
THPPTK-PEG | Thioketal | Photo/hypoxia | Nanoparticle | AQ4N | 78.79% after 4 h with a 660 nm laser | 4T1 xenograft breast cancer model | 40 |
Nano TKPP-COF | Dithioketal | Photo | Nanoparticle | DC_AC50 | 97.7% in 12 h with a 660 nm laser | HT-1080 xenograft mouse model | 133 |
Fig. 17 pH-responsive COFs. (A) I. Schematic diagram of chemical structure of DT-COF. II. Schematic representation of the loading of carboplatin. III. Drug release profiles of carboplatin-loaded DT-COF. IV. Formation of a hydrogen bond between carboplatin and DT-COF from DFT calculation. Reproduced from ref. 62 with permission from Microporous and Mesoporous Materials, Copyright 2020. (B) I. Schematic representation for preparing JUC-556-[HZ]X with E/Z isomerization. II. Release of Ara-C from the channels of JUC-556-[HZ]X in acidic solution. III. Drug release profiles and IV. reversibility of Ara-C-loaded JUC-556-[HZ]0.50 in a simulated cancer fluid. Reproduced from ref. 71 with permission from Small, Copyright 2021. |
Fang's group constructed two 3D pH-triggered COFs based on a hydrazone derivative with E/Z interconversion, with the abbreviations of JUC-556-[HZ]x (Z) and JUC-556-[HZ]x (E), where x was the proportion of one of the linkers HZ (x = 0.25, 0.50, 0.75, or 1.00).71 These functionalized COFs were used for loading cytarabine (Ara-C) in a pH-responsive way. Specifically, the addition of acid (trifluoroacetic acid, TFA) or base (triethylamine, Et3N) induced E/Z isomerization in the dissociative HZ units, which could be inspected by UV-vis absorption spectroscopy, showing a characteristic absorption band at 249 nm for Z isomerization and 263 nm for E isomerization. Finally, JUC-556-[HZ]0.50 displayed the best performance in terms of drug release, which reached 74.56% in pH = 4.8 buffer within 72 h, while only 18.59% in pH = 7.4 buffer, as illustrated in Fig. 17(B).
Fig. 18 Redox-responsive COFs. (A) Preparation of drug-loaded F68@SS-COFs and their intracellular GSH-responsive drug release. Reproduced from ref. 64 with permission from Macromolecular Rapid Communications, Copyright 2020. (B) Synthesis of 5-Fu⊂nano DSPP-COF and its treatment application, and the 5-Fu release behavior triggered by GSH with different concentrations. Reproduced from ref. 86 with permission from Chemical Science, Copyright 2023. (C) Preparation of matrine-loaded TPE-ss COF (TPE-ss COF@matrine) and intracellular GSH-responsive drug release. Reproduced from ref. 76 with permission from Small, Copyright 2022. (D) Schematic illustration of DiSe-Por-DOX for combination antitumor therapy and drug release curves with different conditions. Reproduced from ref. 80 with permission from Journal of Materials Chemistry B, Copyright 2022. |
Recently, Dong's group constructed a host–guest supramolecular system (5-Fu⊂nano DSPP-COF) by immersing nano DSPP-COF in a 5-Fu methanol solution at room temperature for 24 h, and the loading content of 5-Fu was calculated to be around 0.86 μmol mg−1.86 The disulfide-involved DSPP-COF showed GSH-triggered biodegradability as expected, which was verified by the experimental results, where up to ∼96.9% of the encapsulated 5-Fu was released from 5-Fu⊂nano DSPP-COF with the external environment of a 10 mM GSH buffered solution, while only ∼11.3% 5-Fu was observed without GSH (Fig. 18(B)). In addition to TME that overexpresses GSH, the intracellular myocyte microenvironment also undergoes overexpression of GSH after ischemia-reperfusion injury. Thus, Huang et al. designed a redox-responsive COF (TPE-ss COF) as an effective drug delivery carrier for matrine, an anti-cryptosporidial drug.76 As shown in Fig. 18(C), the matrine-loaded TPE-ss COF composite (TPE-ss COF@matrine) exhibited an evident GSH-dependent release profile, where over 90% of loaded drugs was released from the COF within 80 h in response to the condition of pH = 5.0 and 10 mM GSH, while only about 30% or 10% release could be seen under pH = 5.0 or 7.4 in the same period.
Interestingly, Lou et al. prepared a diselenium-bridged COF (DiSe-Por-DOX) for the effective encapsulation and highly controlled release of DOX.80 When internalization in tumor cells, the disrupting of diselenium bonds (Se–Se) in the high GSH tumor internal environment could not only promote the release of DOX, but also boost the generation of intracellular ROS, which broke the redox homeostasis of cancer cells (Fig. 18(D)). As expected, the release of DOX increased with a decrease in acidity (pH = 7.4/6.5/5.5) or increase in the GSH concentration (0.1 mM to 10 mM) or combined condition.
Fig. 19 Hypoxia-responsive COFs. (I) Process for the synthesis of TA-COF-P@CT for light-activated hypoxia-sensitive cancer treatment. Hydrodynamic size distribution and TEM images (inset) of (II) TB-COF-P@CT and (III) TA-COF-P@CT before and after the treatment with Na2S2O4. Reproduced from ref. 68 with permission from Nano Letters, Copyright 2021. |
Fig. 20 Protein-responsive COFs. I. Process for the synthesis of PER@PDA-COF-1 and its structure. II. Donor–acceptor based strategy for the loading of bio-actives and albumin-triggered cellular release. III. Schematic representation of albumin-triggered mitoxantrone (MXT) release from the porous host PER@PDA-COF-1. IV. DLS profile of PER@PDA-COF-1 in the presence of different concentrations of albumin. V. Time-dependent drug release profiles of MXT-PER@PDA-COF-1 alone and HAS-MXT-PER@PDA-COF-1 nano-formulations with different albumin concentrations. Reproduced from ref. 79 with permission from Chemical Science, Copyright 2022. |
Fig. 21 Photo-responsive COFs. (A) Process for the synthesis of AQ4N@THPPTK-PEG NPs and their combined chemotherapy and photodynamic therapy effects and photo-responsive release behavior under 660 nm laser irradiation. Reproduced from ref. 77 with Acta Biomaterialia, Copyright 2022. (B) Preparation of DC_AC50@nano TKPP-COF and its “on–off” photo-responsive release behavior under 660 nm laser irradiation. Reproduced from ref. 133 with Chemical Communications, Copyright 2023. |
A variety of stimuli-responsive COFs that can be activated by diverse stimuli including pH, GSH, hypoxia, albumin and light was discussed in this review. For example, imine-based COFs condensed through the Schiff-base reaction have attracted the most attention for pH-responsive release due to the pH-sensitivity of their imine bonds. In addition, by introducing S–S or Se–Se bonds, the spontaneous redox reactions between materials and the microenvironment make GSH-response possible. Similarly, the hypoxia-response is achievable through the introduction of azo bonds in COFs. Furthermore, in the case of 2D COFs for DDS, the layer-by layer exfoliating strategy seems to be a powerful method for controlled release when using a specific protein exfoliating agent.
Based on the current research status, there are still many challenges associated with COFs in bio-applications. Firstly, from the perspective of COF synthesis, more efforts should devoted to obtaining high-quality nanoscale COFs, especially for high crystallinity and porosity. The lack of finite porosity and functionality may result in potential flaws. Secondly, COFs need to be stable enough to carry their entrapped cargos to the biological targets to avoid trigger leakage, causing toxicological effects. Thirdly, there is little or no research on the in vitro or in vivo pharmacokinetics of COFs. The investigation of possible cumulative and chronic toxicity of monomers degraded from COFs has not been formulated. Fourthly, there is indeed a long way to go before translating basic research results into clinical practice.
In conclusion, the library of COFs that are suitable for the integration and delivery of cargos is growing as research proceeds, and it will continue to be an attractive and challenging research field. We hope that this review article will inspire follow-up studies on COFs for the integration, delivery and controlled release of bioactives.
This journal is © The Royal Society of Chemistry 2024 |