Abstract

•Fully sp2-C-conjugated TPE-based isomeric POPs are reported for CEES detoxification•Porous photosensitizers offer ROS generation even under visible light and ultrasound•Catalytic oxidation of CEES is achieved under ambient atmosphere without toxic products•The composition of the generated ROS types is engineered by the isomeric POP structures Selective oxidation of sulfide to sulfoxide, controlling overoxidized toxic products, is a highly demanding strategy for the detoxification of sulfur mustard simulants (CEES). Singlet oxygen is a widely studied mild oxidant, but it is utilized mainly in an oxygen-abundant environment for CEES oxidation rather than atmospheric ambient conditions. Herein, we present the isomeric TPE-based porous organic polymers (POPs) to photo-oxidize CEES even under ambient conditions without toxic byproducts. Furthermore, we also discovered that superoxide and hydroxyl radicals participate in the oxidative mechanism to a greater extent in ambient conditions than in the O2 atmosphere. In addition, we demonstrate the sono-catalytic selective oxidation of CEES by POPs by ultrasound irradiation for the first time. Our findings will give inspiration for further prospective applications of POPs as photo- and sono-catalysts. The development of efficient strategies for the sustainable detoxification of mustard gas simulants has longstanding demand for human safety. Here, we present for the first time a photo- and sono-catalyzed selective detoxification of mustard gas simulants under ambient conditions using conveniently prepared porous organic polymer (POP) catalysts. We developed a microwave-assisted synthesis of three isomeric tetraphenylethylene-based POPs (TPo, TPm, and TPp) bearing fully sp2-hybridized carbon frameworks. Among the three isomers, TPm efficiently generated 1O2, whereas TPp generated both O2.– and HO. under visible light irradiation, both TPm and TPp efficiently generated ROS to selectively convert 2-chloroethyl ethyl sulfide (CEES) into nontoxic 2-chloroethyl ethyl sulfoxide (CEESO) in the atmospheric conditions, through a known conversion mechanism. TPm and TPp can also generate 1O2 under ultrasound irradiation. This work provides insight into designing new POP photo- and sono-catalysts to generate ROS in widespread applications. The development of efficient strategies for the sustainable detoxification of mustard gas simulants has longstanding demand for human safety. Here, we present for the first time a photo- and sono-catalyzed selective detoxification of mustard gas simulants under ambient conditions using conveniently prepared porous organic polymer (POP) catalysts. We developed a microwave-assisted synthesis of three isomeric tetraphenylethylene-based POPs (TPo, TPm, and TPp) bearing fully sp2-hybridized carbon frameworks. Among the three isomers, TPm efficiently generated 1O2, whereas TPp generated both O2.– and HO. under visible light irradiation, both TPm and TPp efficiently generated ROS to selectively convert 2-chloroethyl ethyl sulfide (CEES) into nontoxic 2-chloroethyl ethyl sulfoxide (CEESO) in the atmospheric conditions, through a known conversion mechanism. TPm and TPp can also generate 1O2 under ultrasound irradiation. This work provides insight into designing new POP photo- and sono-catalysts to generate ROS in widespread applications. IntroductionThe dangerous chemical warfare agents (CWAs), such as phosgene, lewisite, and other hazardous substances, have been a severe threat to humans since their initial use in World War I. These chemicals include a notorious sulfur mustard, 1,2-bis(2-chloroethyl)sulfide (military designation: HD), which has resulted in millions of casualties due to its low cost and convenient production.1Ganesan K. Raza S.K. Vijayaraghavan R. Chemical warfare agents.J. Pharm. Bioallied Sci. 2010; 2: 166-178https://doi.org/10.4103/0975-7406.68498Crossref PubMed Google Scholar Consequently, finding a way to degrade these CWAs safely has been urgently demanded.2Oheix E. Gravel E. Doris E. Catalytic processes for the neutralization of sulfur mustard.Chem. Eur. J. 2021; 27: 54-68https://doi.org/10.1002/chem.202003665Crossref PubMed Scopus (16) Google Scholar Commonly used detoxification methods, including dehydrohalogenation,3Wagner G.W. Koper O.B. Lucas E. Decker S. Klabunde K.J. Reactions of VX, GD, and HD with nanosize CaO: autocatalytic dehydrohalogenation of HD.J. Phys. Chem. B. 2000; 104: 5118-5123https://doi.org/10.1021/jp000101jCrossref Scopus (283) Google Scholar hydrolysis,4Irvine D.A. Earley J.P. Cassidy D.P. Harvey S.P. Biodegradation of sulfur mustard hydrolysate in the sequencing batch reactor.Water Sci. Technol. 1997; 35: 67-74https://doi.org/10.2166/wst.1997.0014Crossref Google Scholar and oxidation,5Wagner G.W. Procell L.R. Yang Y.-C. Bunton C.A. Molybdate/peroxide oxidation of mustard in microemulsions.Langmuir. 2001; 17: 4809-4811https://doi.org/10.1021/la010334hCrossref Scopus (76) Google Scholar,6Ringenbach C.R. Livingston S.R. Kumar D. Landry C.C. Vanadium-doped acid-prepared mesoporous silica: synthesis, characterization, and catalytic studies on the oxidation of a mustard gas analogue.Chem. Mater. 2005; 17: 5580-5586https://doi.org/10.1021/cm051372fCrossref Scopus (48) Google Scholar aim to transform the structure to inhibit the formation of a reactive cationic sulfonium ring. The first two methods are limited by slow rates due to the immiscibility of HD in water, whereas oxidation is kinetically favorable, as it can be performed in organic solvents. The primary requirement in the oxidation process is selective oxidation of sulfur in CEES to the formation of nontoxic CEESO, because the overoxidized product, 2-chloroethyl ethyl sulfone (CEESO2), is known to be toxic. Among oxidants, singlet oxygen (1O2) is a promising route to produce nontoxic products, as metal oxides and peroxides partially generate the undesired toxic sulfone products.7Atilgan A. Cetin M.M. Yu J. Beldjoudi Y. Liu J. Stern C.L. Cetin F.M. Islamoglu T. Farha O.K. Deria P. Post-synthetically elaborated BODIPY-based porous organic polymers (POPs) for the photochemical detoxification of a sulfur mustard simulant.J. Am. Chem. Soc. 2020; 142: 18554-18564https://doi.org/10.1021/jacs.0c07784Crossref PubMed Scopus (41) Google Scholar, 8Liu Y. Howarth A.J. Hupp J.T. Farha O.K. Selective photooxidation of a mustard-gas simulant catalyzed by a porphyrinic metal–organic framework.Angew. Chem. In. Ed. 2015; 54: 9001-9005https://doi.org/10.1002/anie.201503741Crossref PubMed Scopus (203) Google Scholar, 9Wang H. Wagner G.W. Lu A.X. Nguyen D.L. Buchanan J.H. McNutt P.M. Karwacki C.J. Photocatalytic oxidation of sulfur mustard and its simulant on BODIPY-incorporated polymer coatings and fabrics.ACS Appl. Mater. Interfaces. 2018; 10: 18771-18777https://doi.org/10.1021/acsami.8b04576Crossref PubMed Scopus (39) Google Scholar Therefore, 1O2 generation strategy is most desirable for selective HD detoxification to nontoxic products without overoxidation.Recently, some photosensitizers (PSs) have been developed with the focus on the generation of their reactive oxygen species (ROS) in aqueous conditions.10Zhang T. Xing G. Chen W. Chen L. Porous organic polymers: a promising platform for efficient photocatalysis.Mater. Chem. Front. 2020; 4: 332-353https://doi.org/10.1039/C9QM00633HCrossref Google Scholar Still, generation of ROS by PSs for the detoxification of HD needs to be re-evaluated in organic media because of its water insolubility.11Liu Y. Buru C.T. Howarth A.J. Mahle J.J. Buchanan J.H. DeCoste J.B. Hupp J.T. Farha O.K. Efficient and selective oxidation of sulfur mustard using singlet oxygen generated by a pyrene-based metal–organic framework.J. Mater. Chem. A. 2016; 4: 13809-13813https://doi.org/10.1039/C6TA05903ACrossref PubMed Google Scholar,12McNutt P.M. Kelly K.E. Altvater A.C. Nelson M.R. Lyman M.E. O’Brien S. Conroy M.T. Ondeck C.A. Bodt S.M.L. Wolfe et al.Dose-dependent emergence of acute and recurrent corneal lesions in sulfur mustard-exposed rabbit eyes.Toxicol. Lett. 2021; 341: 33-42https://doi.org/10.1016/j.toxlet.2021.01.016Crossref PubMed Scopus (6) Google Scholar In this context, the development of a new PS capable of efficiently generating ROS under light irradiation in organic media is currently a topic of intense research for the safe and effective catalytic oxidation of HD.13Long Z.-H. Luo D. Wu K. Chen Z.-Y. Wu M.-M. Zhou X.-P. Li D. Superoxide ion and singlet oxygen photogenerated by metalloporphyrin-based metal-organic frameworks for highly efficient and selective photooxidation of a sulfur mustard simulant.ACS Appl. Mater. Interfaces. 2021; 13: 37102-37110https://doi.org/10.1021/acsami.1c08840Crossref PubMed Scopus (12) Google Scholar, 14Jia P.-P. Xu L. Hu Y.-X. Li W.-J. Wang X.-Q. Ling Q.-H. Shi X. Yin G.-Q. Li X. Sun H. et al.Orthogonal self-assembly of a two-step fluorescence-resonance energy transfer system with improved photosensitization efficiency and photooxidation activity.J. Am. Chem. Soc. 2021; 143: 399-408https://doi.org/10.1021/jacs.0c11370Crossref PubMed Scopus (41) Google Scholar, 15Li W.-J. Wang X.-Q. Zhang D.-Y. Hu Y.-X. Xu W.-T. Xu L. Wang W. Yang H.-B. Artificial light-harvesting systems based on AIEgen-branched rotaxane dendrimers for efficient photocatalysis.Angew. Chem. Int. Ed. 2021; 60: 18761-18768https://doi.org/10.1002/anie.202106035Crossref PubMed Scopus (35) Google Scholar, 16Son F.A. Wasson M.C. Islamoglu T. Chen Z. Gong X. Hanna S.L. Lyu J. Wang X. Idrees K.B. Mahle J.J. et al.Uncovering the role of metal–organic framework topology on the capture and reactivity of chemical warfare agents.Chem. Mater. 2020; 32: 4609-4617https://doi.org/10.1021/acs.chemmater.0c00986Crossref Scopus (36) Google Scholar, 17Xia S.-G. Zhang Z. Wu J.-N. Wang Y. Sun M.-J. Cui Y. Zhao C.-L. Zhong J.-Y. Cao W. Wang H. et al.Cobalt carbide nanosheets as effective catalysts toward photothermal degradation of mustard-gas simulants under solar light.Appl. Catal. B Environ. 2021; 284: 119703https://doi.org/10.1016/j.apcatb.2020.119703Crossref Scopus (12) Google Scholar Furthermore, we found that ultrasound (US) is also a powerful tool, with the advantage of deep penetration, and can be used for the catalytic detoxification of a sulfur mustard simulant, which, however, has never been exploited.18Lin X. Song J. Chen X. Yang H. Ultrasound-activated sensitizers and applications.Angew. Chem. Int. Ed. 2020; 59: 14212-14233https://doi.org/10.1002/anie.201906823Crossref PubMed Scopus (126) Google Scholar,19Son S. Kim J.H. Wang X. Zhang C. Yoon S.A. Shin J. Sharma A. Lee M.H. Cheng L. Wu J. et al.Multifunctional sonosensitizers in sonodynamic cancer therapy.Chem. Soc. Rev. 2020; 49: 3244-3261https://doi.org/10.1039/C9CS00648FCrossref PubMed Google ScholarSeveral metal-organic framework (MOF) photosensitizers with potent properties, such as structural diversity, large surface area, and tunable functionality have been recently reported.20Vellingiri K. Philip L. Kim K.-H. Metal–organic frameworks as media for the catalytic degradation of chemical warfare agents.Coord. Chem. Rev. 2017; 353: 159-179https://doi.org/10.1016/j.ccr.2017.10.010Crossref Scopus (67) Google Scholar While MOF-based photosensitizers exhibit good photo-oxidation performance, their structural diversity is limited because they are mainly composed of a few photosensitizing modalities, such as porphyrin and BODIPY with heavy atoms.11Liu Y. Buru C.T. Howarth A.J. Mahle J.J. Buchanan J.H. DeCoste J.B. Hupp J.T. Farha O.K. Efficient and selective oxidation of sulfur mustard using singlet oxygen generated by a pyrene-based metal–organic framework.J. Mater. Chem. A. 2016; 4: 13809-13813https://doi.org/10.1039/C6TA05903ACrossref PubMed Google Scholar,21Liu Y. Moon S.-Y. Hupp J.T. Farha O.K. Dual-function metal–organic framework as a versatile catalyst for detoxifying chemical warfare agent simulants.ACS Nano. 2015; 9: 12358-12364https://doi.org/10.1021/acsnano.5b05660Crossref PubMed Scopus (174) Google Scholar, 22Lee D.T. Jamir J.D. Peterson G.W. Parsons G.N. Protective fabrics: metal-organic framework textiles for rapid photocatalytic sulfur mustard simulant detoxification.Matter. 2020; 2: 404-415https://doi.org/10.1016/j.matt.2019.11.005Abstract Full Text Full Text PDF Scopus (54) Google Scholar, 23Atilgan A. Islamoglu T. Howarth A.J. Hupp J.T. Farha O.K. Detoxification of a sulfur mustard simulant using a BODIPY-functionalized zirconium-based metal–organic framework.ACS Appl. Mater. Interfaces. 2017; 9: 24555-24560https://doi.org/10.1021/acsami.7b05494Crossref PubMed Scopus (81) Google Scholar In addition, since most of the relevant experiments have been performed in an external O2 atmosphere, not in ambient conditions, the catalytic performance for practical applications has received less attention. Moreover, small-molecule photocatalysts that enable oxidation at atmospheric O2 concentrations (∼21%) have been reported, but only under very high pressure (∼9 bar). Therefore, it is essential to develop practical strategies to ensure human health and healthy ecosystems.24Emmanuel N. Bianchi P. Legros J. Monbaliu J.-C.M. A safe and compact flow platform for the neutralization of a mustard gas simulant with air and light.Green. Chem. 2020; 22: 4105-4115https://doi.org/10.1039/D0GC01142HCrossref Google ScholarPorous organic polymers (POPs) have emerged as attractive photocatalytic materials owing to their modular structure, inherent porosity, and long-term stability.10Zhang T. Xing G. Chen W. Chen L. Porous organic polymers: a promising platform for efficient photocatalysis.Mater. Chem. Front. 2020; 4: 332-353https://doi.org/10.1039/C9QM00633HCrossref Google Scholar A fully sp2-hybridized carbon-conjugated system has been an attractive platform for developing a new photosensitizer, as this system can harvest a broad range of visible light by narrowing the band gap.25Jin E. Asada M. Xu Q. Dalapati S. Addicoat M.A. Brady M.A. Xu H. Nakamura T. Heine T. Chen Q. et al.Two-dimensional sp2 carbon–conjugated covalent organic frameworks.Science. 2017; 357: 673-676https://doi.org/10.1126/science.aan0202Crossref PubMed Scopus (566) Google Scholar,26Liu S. Feng G. Tang B.Z. Liu B. Recent advances of AIE light-up probes for photodynamic therapy.Chem. Sci. 2021; 12: 6488-6506https://doi.org/10.1039/D1SC00045DCrossref PubMed Google Scholar The π-networks also allow exciton migration and prevent backward charge recombination.27Jin E. Lan Z. Jiang Q. Geng K. Li G. Wang X. Jiang D. 2D sp2 carbon-conjugated covalent organic frameworks for photocatalytic hydrogen production from water.Chem. 2019; 5: 1632-1647https://doi.org/10.1016/j.chempr.2019.04.015Abstract Full Text Full Text PDF Scopus (250) Google Scholar In addition, the large surface areas and pore volumes of POPs can be expected to accelerate ROS generation owing to the amount of O2 gas that they encapsulate.28Wei G. Huang L. Shen Y. Huang Z. Xu X. Zhao C. Porphyrin-based porous organic frameworks as oxygen reservoirs to overcome tumor hypoxia for enhanced photodynamic therapy.Adv. Ther. 2019; 2: 1900059https://doi.org/10.1002/adtp.201900059Crossref Scopus (12) Google ScholarIn this context, for the first time, we have designed POP-based photo/sonosensitizers using tetraphenylethylene (TPE), a well-known PS modality, for CEES detoxification. We prepared three TPE-based isomeric polymers, TPo, TPm, and TPp, using ortho-, meta-, and para-xylylene dicyanide, respectively (Figure 1). We observed pronounced performance in CEES oxidation by TPm and TPp in a protic solvent, with significant conversion rates under both O2 and ambient atmospheres. We herein also propose mechanisms for CEES detoxification using POPs by reaction with 1O2 and other ROS.Results and discussionThree two-dimensional POPs (TPo, TPm, and TPp) were synthesized via the well-known Knoevenagel condensation process, the reaction of 1,1,2,2-tetrakis(4-formylphenyl)ethene (TPE-4CHO) with xylylene dicyanide isomers (ortho-, meta-, and para-) in the presence of base produced fully sp2-conjugated porous polymers.25Jin E. Asada M. Xu Q. Dalapati S. Addicoat M.A. Brady M.A. Xu H. Nakamura T. Heine T. Chen Q. et al.Two-dimensional sp2 carbon–conjugated covalent organic frameworks.Science. 2017; 357: 673-676https://doi.org/10.1126/science.aan0202Crossref PubMed Scopus (566) Google Scholar,29Yassin A. Trunk M. Czerny F. Fayon P. Trewin A. Schmidt J. Thomas A. Structure-thermodynamic-property relationships in cyanovinyl-based microporous polymer networks for the future design of advanced carbon capture materials.Adv. Funct. Mater. 2017; 27: 1700233https://doi.org/10.1002/adfm.201700233Crossref Scopus (28) Google Scholar We have utilized a much faster microwave-assisted Knoevenagel condensation reaction for the first time. The reaction time was significantly reduced, from 3 days with a conventional solvothermal method to 3 h, ascribable to homogeneous heating and rapid nucleation under microwave irradiation.30Kang D.W. Lim K.S. Lee K.J. Lee J.H. Lee W.R. Song J.H. Yeom K.H. Kim J.Y. Hong C.S. Cost-effective, high-performance porous-organic-polymer conductors functionalized with sulfonic acid groups by direct postsynthetic substitution.Angew. Chem. Int. Ed. 2016; 55: 16123-16126https://doi.org/10.1002/anie.201609049Crossref PubMed Scopus (57) Google ScholarWe anticipated that TPE and xylylene dicyanide would serve as electron donor and electron acceptor, respectively, to construct a porous donor-acceptor framework. The nitrile unit, as a strong electron-withdrawing group, lowers the lowest unoccupied molecular orbital in the sp2-conjugated system, enabling the suppression of charge recombination.31Chen W. Wang L. Mo D. He F. Wen Z. Wu X. Xu H. Chen L. Modulating benzothiadiazole-based covalent organic frameworks via halogenation for enhanced photocatalytic water splitting.Angew. Chem. Int. Ed. 2020; 59: 16902-16909https://doi.org/10.1002/anie.202006925Crossref PubMed Scopus (144) Google Scholar For this reason, the synthesized POPs can be expected to have great potential as promising PSs.Powder X-ray diffraction data indicate that the polymeric materials we prepared are amorphous (Figure S1). Following the Knoevenagel condensation, an intensity of C=O stretching peak in the infrared (IR) spectra centered at 1,693 cm−1 for the aldehyde in TPE-4CHO decreased, implying the presence of extended sp2-C conjugation forming a porous network structure (Figures S2–S4). Interestingly, the distinct C≡N stretching peaks at 2,250 cm−1 in all the xylylene dicyanide isomers were red-shifted to 2,216 cm−1 (TPm and TPp) or 2,197 cm−1 (TPo) (Figure 2A). This red-shift is attributed to the weakening of the C–N bond strength resulting from the condensation reaction. X-ray photoelectron spectroscopy analysis suggested that the elements C, N, and O were present in all the POPs, as shown in Figures S5 and S6. As the chemical environment of the nitrile group in each POP isomer is different, the binding energies of the N1s electrons were seen to be different from each other. Solid-state 13C nuclear magnetic resonance (13C ssNMR) data were recorded to obtain detailed structural information on the POPs (Figure 2B). In all the NMR spectra, peaks (region 1) were observed at approximately 170 ppm, assigned to unreacted aldehyde. The peak due to unreacted aldehyde was more intense in TPp, suggesting a lower degree of polymerization than in the other two POPs. Moreover, sp2 (region 2, phenyl ring) and sp (region 3, nitrile group) carbon peaks were observed at 150–130 and 110 ppm, respectively. Peak positioned in region 4 (below 50 ppm) indicates residual n-hexane and sp3 carbons in unreacted benzyl cyanide in the polymer backbone.Figure 2Characterization data of TPo, TPm, and TPpShow full caption(A) IR spectra of TPo, TPm, and TPp.(B) Solid-state 13C NMR spectra.(C and D) UV-vis spectra in the solid-state and dispersed in hexane, respectively.(E) Band structure diagram, the top and bottom bars indicate the conduction and valence bands of the POPs, respectively.(F) N2 isotherms at 77 K. Filled and open symbols indicate adsorption and desorption, respectively.View Large Image Figure ViewerDownload Hi-res image Download (PPT)Furthermore, we investigated the optical and electronic properties of the POPs to verify their potential as photosensitizers. In the solid-state UV-vis absorption spectra (Figure 2C), TPo shows a broad range of absorption in the visible region, whereas both TPm and TPp display a maximum absorption peak at 418 nm with a 200–600 nm absorption range. In addition, the photoluminescence quantum yield (PLQY) was determined to be 0% (0%) for TPo, 11% (13%) for TPm, and 13% (15%) for TPp at 400 (450) nm excitation (Figure S7). However, when the POPs were dispersed in n-hexane to measure the absorption spectra, they showed a disparity among the POP isomers (Figure 2D). The maximum absorption peaks were shifted to 488 nm (TPp) and 457 nm (TPm), and no specific peak for TPo was observed. To understand the difference in optical response, we measured the particle size of POP isomers by scanning electron microscopy (Figure S8). The POPs showed diverse particle sizes (TPo: ∼900 nm, TPm: ∼50 nm, TPp: ∼500 nm), despite using the same synthetic method. Although it is difficult to obtain accurate absorbance data of TPo in solution because of its low dispersity, this corresponds to the fact that the absorption peaks of the nanoparticles could be dependent on their particle size.32Dai C. Yang D. Zhang W. Bao B. Cheng Y. Wang L. Far-red/near-infrared fluorescent conjugated polymer nanoparticles with size-dependent chirality and cell imaging applications.Polym. Chem. 2015; 6: 3962-3969https://doi.org/10.1039/C5PY00344JCrossref Google Scholar The band gaps of TPo, TPm, and TPp, calculated by Tauc plots from ssUV data, were 1.74, 2.32, and 2.31 eV, respectively (Figure S9). In addition, we used cyclic voltammetry with Ag/AgCl as a reference electrode to determine the potential levels of the POPs (Figure S10). For the oxidation process, the potential levels of the HOMOs for POPs were calculated as −5.77 eV (TPo), −5.58 eV (TPm), and −5.63 eV (TPp) versus vacuum level, respectively. By combining all the analytical data, we have profiled band diagrams of the POPs (Figure 2E) to visually identify their distinct optical and electronic properties.We next explored the thermal stability of the polymers prior to gas isotherm measurements and found that they were structurally stable up to 150°C (Figure S11). The intrinsic porosity of the POPs was investigated by measuring nitrogen isotherms at 77 K after degassing at 120°C for 12 h (Figures 2F and S12). The Brunauer-Emmett-Teller surface areas were calculated to be 8.8 m2 g−1 (TPo), 113 m2 g−1 (TPm), and 79 m2 g−1 (TPp). The main pore sizes were found to be 1.18 nm (TPm) and 1.48 nm (TPp) by pore size distribution analysis using the N2-DFT model, except for the non-porous TPo isomer. Taking all the analysis results into consideration, we propose the two-dimensional structures for TPm and TPp shown in Figure S13.The ROS generation abilities of TPE-4CHO and TPp were evaluated in a methanol solution with 9,10-anthracenediyl-bis(methylene)dimalonic acid (ABDA) as an indicator of 1O2 under white LED lamp irradiation (Figures 3A and S14). As the TPE monomer is a representative aggregation-induced emission luminogen (AIEgen), TPE-4CHO displays no sensitizing performance in organic solvents because it cannot aggregate.33Hu F. Xu S. Liu B. Photosensitizers with aggregation-induced emission: materials and biomedical applications.Adv. Mater. 2018; 30: 1801350https://doi.org/10.1002/adma.201801350Crossref PubMed Scopus (423) Google Scholar,34Ni J. Wang Y. Zhang H. Sun J.Z. Tang B.Z. Aggregation-induced generation of reactive oxygen species: mechanism and photosensitizer construction.Molecules. 2021; 26: 268https://doi.org/10.3390/molecules26020268Crossref Scopus (24) Google Scholar However, TPp generates an appreciable amount of 1O2 under the same solvent conditions as the monomer. It was previously known that TPE moieties with quadrilateral cores forming a porous framework restrict intramolecular rotation, vibration, and motion, revealing similar performance to AIE-based photosensitizers.35Shustova N.B. McCarthy B.D. Dinca M. Turn-on fluorescence in tetraphenylethylene-based metal–organic frameworks: an alternative to aggregation-induced emission.J. Am. Chem. Soc. 2011; 133: 20126-20129https://doi.org/10.1021/ja209327qCrossref PubMed Scopus (532) Google Scholar,36Zhang R. Feng G. Zhang C.-J. Cai X. Cheng X. Liu B. Real-time specific light-up sensing of transferrin receptor: image-guided photodynamic ablation of cancer cells through controlled cytomembrane disintegration.Anal. Chem. 2016; 88: 4841-4848https://doi.org/10.1021/acs.analchem.6b00524Crossref PubMed Scopus (44) Google Scholar Furthermore, the polymerization was found to facilitate the intersystem crossing (ISC) process by dense energy-level distributions compared with the monomeric molecule.37Wu W. Mao D. Xu S. Kenry Hu F. Li X. Kong D. Liu B. Polymerization-enhanced photosensitization.Chem. 2018; 4: 1937-1951https://doi.org/10.1016/j.chempr.2018.06.003Abstract Full Text Full Text PDF Scopus (150) Google Scholar The photo-induced ROS generation of each TPo, TPm, and TPp was then examined under the conditions shown in Figure 2A but using 405 nm light (Figures 3B and S15). Among the three isomers, TPm exhibited the most efficient generation of 1O2, presumably from faster ROS diffusion due to the larger surface area, smaller particle size, and more overlapped energy states than in the other two isomers.38Lan G. Ni K. Veroneau S.S. Luo T. You E. Lin W. Nanoscale metal–organic framework hierarchically combines high-Z components for multifarious radio-enhancement.J. Am. Chem. Soc. 2019; 141: 6859-6863https://doi.org/10.1021/jacs.9b03029Crossref PubMed Scopus (47) Google Scholar We also examined the 1O2 generation ability of the POPs using a mixture of p-nitroso dimethylaniline (RNO) and imidazole, an alternative 1O2 indicator, and found that TPm showed a marked decrease in RNO absorbance at 420 nm (Figure S16).39Herman J. Neal S.L. Efficiency comparison of the imidazole plus RNO method for singlet oxygen detection in biorelevant solvents.Anal. Bioanal. Chem. 2019; 411: 5287-5296https://doi.org/10.1007/s00216-019-01910-2Crossref PubMed Scopus (10) Google Scholar Next, as superoxide (O2.–) generation could be predicted from the band gap diagram (Figure 2E), we used dihydrorhodamine123 (DHR123) as a peroxide indicator. The absorbance peak of DHR123 at approximately 500 nm selectively increased in the case of TPp over TPo and TPm (Figures 3C and S17). Moreover, TPp was found to generate not only O2.– but numerous hydroxyl radicals (HO∙) in the RNO experiment.40Kraljić I. Trumbore C.N. p-Nitrosodimethylaniline as an OH radical scavenger in radiation chemistry.J. Am. Chem. Soc. 1965; 87: 2547-2550https://doi.org/10.1021/ja01090a004Crossref Scopus (156) Google Scholar We speculate that HO∙ is produced from generated O2.– (Figures 3D and S18).41Parrino F. Livraghi S. Giamello E. Ceccato R. Palmisano L. Role of hydroxyl, superoxide, and nitrate radicals on the fate of bromide ions in photocatalytic TiO2 suspensions.ACS Catal. 2020; 10: 7922-7931https://doi.org/10.1021/acscatal.0c02010Crossref Scopus (26) Google Scholar,42Bienert G.P. Schjoerring J.K. Jahn T.P. Membrane transport of hydrogen peroxide.Biochim. Biophys. Acta Biomembr. 2006; 1758: 994-1003https://doi.org/10.1016/j.bbamem.2006.02.015Crossref PubMed Scopus (788) Google Scholar Consequently, it is notable that TPp can efficiently generate both 1O2 and HO∙.To further validate the type of ROS generated by POPs under light irradiation, electron paramagnetic resonance (EPR) data were collected. 1O2 generation by POPs was confirmed using tetramethyl-4-piperidone (TEMP) as a spin-trapping agent (Figure 3E). The intensity of TEMP with TPm was recorded to be the highest compared with other cases, which agrees well with the observations in the absorbance spectra using ABDA and RNO + imidazole indicators (Figures S15 and S16). Moreover, we employed 5,5-dimethyl-1-pyrroline N-oxide (DMPO) as a radical spin trap to evaluate the existence of other radical species. Strong signals assignable to O2.– were found in the EPR spectra, with TPp exhibiting the highest intensity of DMPO (Figure 3F), consistent with the DHR123 absorbance data (Figure S17).43Gao Z. Lai Y. Tao Y. Xiao L. Li Z. Zhang L. Sun L. Luo F. Creating and tailoring ultrathin two-dimensional uranyl-organic framework nanosheets for boosting photocatalytic oxidation reactions.Appl. Catal. B: Environ. 2021; 297: 120485https://doi.org/10.1016/j.apcatb.2021.120485Crossref Scopus (7) Google Scholar We propose to explain the reason for less 1O2 catalyzing but more radical generation performance of TPp compared with TPm as follows. In general, the amount and types of generated ROS in specific PS systems can compete with each other when triplet electrons are involved in the ROS generation mechanism (Figure 3G). According to the ROS generation mechanism involving conduction band (CB) reported by Zhang and coworkers,44Wang Z.J. Ghasimi S. Landfester K. Zhang K.A.I. Molecular structural design of conjugated microporous poly (benzooxadiazole) networks for enhanced photocatalytic activity with visible light.Adv. Mater. 2015; 27: 6265-6270https://doi.org/10.1002/adma.201502735Crossref PubMed Scopus (197) Google Scholar,45Zhang K. Kopetzki D. Seeberger P.H. Antonietti M. Vilela F. Surface area control and photocatalytic activity of conjugated microporous poly (benzothiadiazole) networks.Angew. Chem. Int. Ed. 2013; 52:

Full Text
Published version (Free)

Talk to us

Join us for a 30 min session where you can share your feedback and ask us any queries you have

Schedule a call