Giter Club home page Giter Club logo

unix_pongpong's Introduction

Unix_pongponglab

Draft: AI based Keyword density analysis

Draft는 AI와 키워드 밀도 분석을 기반으로 논문(nature communication 한정)을 분석해주는 툴로 사용자가 읽어야 하는 부분에 대한 우선순위 선정에 도움을 주는 웹 서비스 입니다.


🚨 notice) 시간 및 서버 상의 이슈(docker설치 불가)로 로컬에서 돌아가던 Api들이 배포되지 못했습니다. 다만 아래의 부분을 통해 실제 url을 기반으로 논문을 분석하는 과정을 mainflow는 코드상에서 테스트 해볼 수 있습니다!

제작 동기

정보 전달 글을 분석할때 가장 좋은 방법은 핵심어를 기준으로 분석하는 것이라 생각합니다. 하지만 현재까지 빠른 논문 분석을 위해서는 관심 및 목적에 따른 핵심어 설정 후, 검색 기능을 통해 문단 하나하나를 읽어보는 방법밖에 없었습니다. 이에 Draft 는 키워드 검색 뿐만이 아니라 키워드의 등장 빈도에 해당하는 데이터도 시각화하여 사용자에게 제공하여 사용자가 읽어야 하는 부분에 대한 우선순위 선정에 도움을 주고자 합니다.

[구성 flow] figma 통해 확인가능

Main flow test 방법

가상환경 생성 및 필요 requirment 설치

  • 3.8버전도 가능하지만, 3.9버전을 권장합니다.
$ conda create -n <venv_name> python=3.9
$ conda activate <venv_name>
$ pip install -r requirments.txt

fastapi_main에서 실제로 main flow 확인해보기

  • 가상환경을 켜둔 상태로 2개의 터미널을 기반으로 확인하는 것이 좋습니다.
  • api 통신 확인을 위해 httpie를 사용하였습니다.
$ cd fastapi_main
$ python flow.py

[AI 기반 keyword extraction]

  • 분석하고 싶은 논문을 nature communication에서 찾아 url 형태로 입력해줍니다.
  • mainflow 확인의 편의를 위해 article은 article.txt파일로 추가 저장됩니다.([예시결과 / AI 기반 keyword extraction] 참조)
$ cd fastapi_main
$ http POST :8000/input url=https://www.nature.com/articles/s41467-022-29811-6#Abs1

[keyword density data]

  • article.txt에서 자동으로 text는 가져오지만 키워드는 사용자가 3개를 제안해야 합니다. ([예시결과 / keyword density data] 참조)
$ cd fastapi_main
$ http POST :8000/analysis \
    url=https://www.nature.com/articles/s41467-022-29811-6 \
    keywords:='["λem", "λmax", "yield"]' \
    article=@article.txt```

### 예시 결과

**[AI 기반 keyword extraction]**

```json
{
    "article": "title\n\nSolution-processable microporous polymer platform for heterogenization of diverse photoredox catalysts\nAbstract\n\nIn contemporary organic synthesis, substances that access strongly oxidizing and/or reducing states upon irradiation have been exploited to facilitate powerful and unprecedented transformations. However, the implementation of light-driven reactions in large-scale processes remains uncommon, limited by the lack of general technologies for the immobilization, separation, and reuse of these diverse catalysts. Here, we report a new class of photoactive organic polymers that combine the flexibility of small-molecule dyes with the operational advantages and recyclability of solid-phase catalysts. The solubility of these polymers in select non-polar organic solvents supports their facile processing into a wide range of heterogeneous modalities. The active sites, embedded within porous microstructures, display elevated reactivity, further enhanced by the mobility of excited states and charged species within the polymers. The independent tunability of the physical and photochemical properties of these materials affords a convenient, generalizable platform for the metamorphosis of modern photoredox catalysts into active heterogeneous equivalents.\n\nIntroduction\n\nPhotoredox catalysis refers to an energy-conversion process in which the absorption of photons by a substance triggers electron transfer. In the past decade, the application of this phenomenon to organic synthesis has evolved from an academic curiosity to an enabling technology widely adopted in pharmaceutical synthesis1,2,3,4,5,6,7. By coupling photoexcitation with other molecular processes, including cooperative catalytic schemes8,9,10, hundreds of powerful and previously inconceivable organic transformations have now been realized. However, compared to the rapid discovery of new photon-dependent reactions, the implementation of these processes on large scale has seen only limited progress due to the high cost of many photocatalysts, challenges in catalyst removal, and restricted light penetration in traditional batch reactors11. Thus, the conception of new technologies12,13,14,15that generally enhance photoredox catalysis on scale is a pressing goal that has attracted immense research effort.\nPhotoredox catalysts can be divided into two major groups (Fig.1a): homogenous materials, which are soluble in the solvents typically used for photoredox catalysis, and heterogeneous, which operate as insoluble solids. Due to ease of synthesis and tunability, the vast majority of newly developed catalysts belong to the former group, which includes small-molecule organic dyes3, such as perylene diimides (PDIs), and organometallic complexes5, such as Ir(ppy)3. Meanwhile, heterogeneous catalysts have the advantages of facile removal from reaction mixtures and recyclability, which are important from both cost and sustainability perspectives6,11. Well-known examples include two-dimensional (2-D) surface photocatalysts, especially semiconductors16such as mesoporous graphitic carbon nitride (mpg-CN)17, and porous three-dimensional (3-D) constructs, such as metal-organic/covalent organic frameworks (MOFs/COFs)18,19,20. The appealing characteristics of these materials arrive at the expense of diversity and tunability: there is often no rational technique to finely alter their photochemical properties while retaining the bulk physical properties. Importantly, the inability of most heterogeneous catalysts to be cast into films or coatings from liquid or solution phase is a considerable practical limitation, particularly towards their synergistic application with emerging technologies such as continuous-flow synthesis10,11.\naClassification of photoredox catalysts according to solubility and geometric dimension of active surface.bTransforming a broad scope of small-molecule photocatalysts into diverse heterogeneous materials through solution-processable polymers.\nA strategy that allows synthetic chemists to simultaneously take advantage of the respective desirable properties of homogeneous and heterogeneous catalysts would be exceptionally empowering. We proposed that one solution might involve a general method for converting any member of the vast toolbox of soluble photocatalysts into equivalent insoluble materials, while maintaining the original photochemical attributes and activity. However, to our knowledge, no heterogenization approach exists with broad scope both in the kinds of dyes incorporated and in the types of catalytic materials accessible. Although several examples of photocatalyst immobilization on insoluble solids have been demonstrated21, each strategy addresses only a small subset of chromophore structures, and the vast majority of supported catalysts lack solution-processability, severely limiting the operational modes that are available (typically only particles or dispersions).\nHerein, by extending our recently reported Pd-catalyzed polycondensation reaction22, we present a method that, in principle, allows for any organic or organometallic photocatalyst to be heterogenized through copolymerization with porosity-promoting bulky organic monomers (triptycenes and spirobifluorenes)23,24,25. We note that while 2D and 3D extended materials are well represented among heterogeneous photocatalysts, examples of one-dimensional (1D) photocatalysts, in which the active components form chains, are relatively rare26,27. Importantly, our linear-polymer structures, while insoluble in the solvents typically employed for photoredox catalysis, remain highly soluble in a few, select handling solvents (DCM and THF). As a result of this processability, our immobilization technique is not restricted to any particular target mode of heterogeneous catalysis but can conveniently access a wide range of known morphologies, including dispersions, films, coatings, textiles, and magnetic nanoparticles.\nAside from solution-processability and broad generality, our 1D polyether platform offers additional valuable benefits compared to existing heterogeneous photoredox catalysts. First, the solubility of organic polymers enables simple characterization of structure, molecular weight distribution, photocatalyst loading, defect concentration, and batch-to-batch quality variations, all of which are often impractical to analyze for conventional insoluble catalysts. Second, traditional materials relying on non-covalent interactions for dye attachment, or labile covalent bonds such as Si–O or Al–O bonds, can exhibit poor chemical stability21. In contrast, the polymers developed in this work exhibit exceptional thermal stability and inertness toward a variety of chemical conditions. Third, the immobilization of small-molecule catalysts on heterogeneous supports often leads to drastically decreased turnover frequencies; however, with our polymer materials, due to a combination of reduced aggregation-induced quenching, highly porous-polymer structure, and energy- and charge-transport processes, very efficient catalysis can be achieved, in some cases with rates exceeding than with the corresponding monomers. Finally, the design of our polymers enables independent control of the physical and photochemical properties without alteration to the synthetic procedure.\n\nResults and discussion\n\nOur recent systematic study22showed that the foundational building blocks of our proposed material,t-Bu-triptycene hydroquinone1-Aand spirobifluorene dibromide1-B, could be cross-coupled to form polymer1with excellent molecular weight, polydispersity, porosity, and film properties (Fig.2a). Through slight alteration of these prior conditions, we were able to introduce chromophore-containing comonomers to generate materials with similar physical attributes as1, but with added photocatalytic capabilities. As a model, we selected the classic dye perylene diimide (PDI) as synthetically versatile redox-active moiety28. With the addition of 10% of PDI dibromide (1-C) in the polycondensation process, we obtained a near-quantitative yield of a deep-red solid (1-PDI) with high molecular weight and moderate polydispersity. The porosity was evaluated through the Brunauer–Emmett–Teller (BET) gas adsorption method, revealing a considerable specific surface area (SSA) of 338 m2/g. The essential role of the triptycene units in endowing the polymer with this exceptional porosity is evident from the very low SSA of related polymer1-PDI-HQ, in whichpara-phenylenes are present instead. Based on the N2adsorption isotherm, the estimated pore size distribution was concentrated in the microporous range (<2 nm, Supplementary Fig.3-1). Furthermore, the polymer showed excellent thermal and chemical stability under controlled conditions (Supplementary Fig.4-2). The extent of incorporation of the PDI was determined by1H NMR analysis to be 10%, in agreement with the fraction of monomer employed during the synthesis.\naSynthesis of1,1-PDI, and1-PDI-HQ. See the Supplementary Information for detailed conditions.bPhotophysical properties.cLight-driven reversible redox process with an amine partner.dKinetics of catalytic photo-oxidation model reaction. BET = Brunauer–Emmett–Teller method.aToo low to be reliably determined from N2adsorption isotherm.bDilute THF solution (1.0 × 10–5–5.0 × 10–7M).cSolid state (powder).\nThe absorption and steady-state photoluminescence spectra of1-PDIcontained distinctive features (λmax= 523 nm, λem= 528 nm in THF solution) characteristic of PDI units (Fig.2b). The solid-state photophysical properties of1-PDIshowed notable differences when compared against monomeric (PDI-Mono) and non-porous (1-PDI-HQ) analogues. Although the molecular dye displayed a very low fluorescence quantum yield in the solid state (QY = 0.02), the chromophores covalently hosted within triptycene-containing frameworks were bright emitters (QY = 0.29 for1-PDI), indicating a reduction in aggregation-based quenching. Monomeric PDIs are known to exhibit a substantial emission red-shift upon aggregation29, and we found that the difference in emission maxima between solution and solid states was substantially smaller for the polymers than for monomers and, in particular, smallest for the porous1-PDI.\nAs a test of the photo-activated reactivity of these polymers, irradiation of a DCM solution of1-PDIwith excess Et3N resulted in isosbestic conversion to a new species, with spectral features indicative of PDI radical anions (Fig.2c)30. Exposure of the resulting solution to atmospheric oxygen regenerated the original polymer over the course of 2 h. Given its promising reversibly stoichiometric reactivity, a dispersion of the catalyst in methanol was evaluated in a simple aerobic photo-oxidation reaction31of methyl phenyl sulfide (2a, Fig.2d). At remarkably low loading of catalyst (0.018 mol%), full consumption of starting material could be observed after 6 h. Parallel experiments at equivalent optical densities of a non-porous analogue (1-PDI-HQ) and a monomeric PDI (PDI-Mono) proceeded at a significantly lower rate and only preceded to 80 and 76% conversion over the same time period. An additional control using a combination ofPDI-Monowith dye-free polyether1(premixed in dichloromethane solution, then dried and dispersed in methanol) showed a reaction rate similar to usingPDI-Monoalone. The reactions did not progress during a 30 min interval with no LED irradiation, confirming that the chosen transformation requires continuous photoirradiation. We hypothesized that, in addition to the steric insulation from aggregation-based quenching, the elevated activity of1-PDIis partially the result of the photoexcited state “hopping” between different dye centers, thereby enhancing the probability of encounter with a substrate or cocatalyst. Indeed, using fluorescence polarization spectroscopy, we observed quencher-dependent behavior consistent with considerable excited-state energy transfer between dye units (for extended discussion, see the Supplementary Information, Section 10.2, Figs.S7-2–S7-4). The mobility of charged states throughout the porous polymer can similarly contribute to an enhanced efficiency. In bulk conductivity measurements, we showed that irradiation of a film of1-PDIin the presence of triethylamine results in an instantaneous and reversible increase in conductivity (Supplementary Fig.7-5), suggesting facile transport of negative charges between PDI units.\nTo illustrate the synthetic versatility of porous linear-polymer catalysts, we applied1-PDIto a representative range of modern photoredox transformations (Fig.3a). On a 1.0 mmol scale, the aerobic oxidation of thioether2ato sulfoxide3awas accomplished in 95% yield using an alcoholic dispersion of1-PDIand irradiation with a 450 nm LED. Photoreduction reactions could be performed through the in situ generation of PDI radical anion under inert atmosphere29. Using triethylamine as the terminal reductant, aryl bromide2bwas successfully hydrodehalogenated. The same catalyst could induce atom-transfer radical addition reactions to alkenes: fluorinated alkyl bromide3cwas prepared in high yield from 5-bromo-1-pentene (2c) and perfluorooctyl iodide32. Since3cis a non-commercial fluorosurfactant essential to other research in our laboratory, we pursued a scaled-up synthesis enabled by catalyst1-PDI. Over 12 g of product could be obtained in a single run, demonstrating the applicability of these porous materials in a preparative setting. Heterogeneous photoredox conditions were also applicable in polar addition to alkenes, as shown by the anti-Markovnikov hydroetherification of2d33. In another activation mode, the oxidative C–H functionalization of electron-rich arene2ewas conducted using potassium bromide as the bromine-atom source17. Both oxidative and reductive modes of radical generation were successful using1-PDIas a photocatalyst, and regioselective addition of these radicals to heteroarenes was accomplished (3fand3g)34,35.\naPerformance of1-PDIacross various reaction classes.bRecyclable usage of1-PDIas a coating on a glass vial in the photo-oxidation of2a.cRecyclable usage of1-PDIdeposited on cotton in the photo-oxidation of2a.dCoated superparamagnetic silica nanoparticles.eContinuous-flow synthesis using a film of1-PDIin glass tubing.\nMany valuable modes of heterogeneous catalysis are accessible due to the solubility profile of poly(arylene ether)s. Using a standard rotary evaporator, a thin film layer could be deposited on the bottom portion of a glass reaction vial, which has maximal exposure to the light source (Fig.3b). The coated vessel could be conveniently reused to perform the photo-oxidation of2a, with negligible change in reaction yield over five cycles. Textile-supported organocatalysts have recently been advanced as convenient promoters for acid- and base-dependent reactions36. We showed that1-PDIcan be robustly deposited onto cotton, and that the resulting fluorescent fabric is also a reusable photo-oxidation catalyst (Fig.3c). The polymer coating could adhere strongly to silica nanoparticles bearing a superparamagnetic iron oxide core (Fig.3d), which facilitates separation from reaction mixtures for which filtration is impractical. The resulting red powder was an effective agent in photoredox reactions, and the application of a handheld permanent magnet allowed for rapid and complete separation of the catalyst from the methanol solution (Supplementary Movie1). The catalytic layer could be instantly desorbed from the surface by immersion in tetrahydrofuran and thus recovered from the silica (Supplementary Movie2).\nContinuous-flow synthesis, referring to processes performed within narrow tubing under constant material input and output, has attracted significant attention as an alternative to traditional batch chemical processes37,38. Flow systems are associated with superior heat exchange, mixing, reproducibility, and safety profiles39. The narrow-channel configuration of the reactors has proved particularly promising in large-scale photocatalysis applications, as enhanced light exposure is afforded with decreased risk of over-irradiation that can induce unwanted side-reactions40. A solution-processable material for the immobilization of diverse photocatalysts along the interior surface of the reactor tubing would provide two main advantages. First, no separation of the dyes from the mobile phase would be required after the reaction. Second, by concentrating the chromophores close to the exterior of the reaction mixture, light exposure could be enhanced. Taking advantage of the excellent film-forming characteristics of our porous polyethers, we showed that flow chemistry could be conducted using a photoredox-active coating as a stationary phase (Fig.3e). To the coolant coil of a spiral reflux condenser, a standard piece of equipment in laboratories that perform microscale organic synthesis, was applied a thin layer of1-PDIas a solution in THF, drying with gentle heating and air flow. With the system under constant irradiation with blue LEDs, reactants for the C–H bromination of2ewere pumped into the coil using a syringe pump (approximate residence time of 60 min). At steady state, a good yield of3ewas measured in the output stream.\nA unique advantage of linear polymers as an immobilization platform is the ability to independently adjust the backbone-derived physical characteristics and the chromophore-derived photochemical attributes (Fig.4a). Achieving adhesion to fluoropolymer surfaces was desirable since inert, flexible substances such as perfluoroalkoxy (PFA) plastics are the most common tubing materials for modern continuous-flow synthesis. Since most poly(arylene ether) polymers such as1-PDIadsorb poorly to fluorinated surfaces, and therefore do not form high-quality films, we wondered if we could enhance the attractive interaction through post-polymerization functionalization of the backbone. We subjected the polymer to C–H fluoroalkylation conditions with photogenerated perfluoroalkyl (n-C17F35) radicals to obtain1-PDI-C17F35with relatively high fluorine content (F/C ratio = 1:4 by XPS). This new polymer was successfully deposited on narrow inner diameter (1/16 in) PFA tubing and effectively employed as a catalytic stationary phase in the flow synthesis of3e(Fig.4b).\naConvenient and independent alteration of catalyst properties.bPost-polymerization functionalization of the backbone enables adhesion to fluorinated surfaces for photoredox catalysis.cStructural variation in the photocatalytic moiety: incorporation of common photoredox chromophores. Numbers in parentheses indicate photoluminescence quantum yields in dilute THF solution (1.0 × 10–5–5.0 × 10–7M).dExample reactions using modified photocatalysts.aMolecular weight distribution measured on1-Bpyprior to Ir incorporation. PMDETA = 1,1,4,7,7-pentamethyldiethylenetriamine.\nThe identity of the covalently-included dye and its doping concentration could be adjusted arbitrarily with little effect on the porosity, solution-processability, and film-forming properties of the polymer. The substitution of other groups in lieu of PDI is synthetically straightforward and requires essentially no procedural modifications (Fig.4c). Perylenes are highly reducing photocatalysts that have been used in applications such as light-activated atom-transfer radical polymerization (photo-ATRP) reactions41. Using perylene-containing1-Peras a catalyst, the polymerization of methyl methacrylate could be performed using sunlight as the energy source. With minimal optimization, the poly(methyl methacrylate) (PMMA,4a) product could be obtained with good molecular weight and polydispersity. Furthermore, the catalytic polymer was easily separated from the product polymer, as the former is strictly insoluble in acetone. The outcome of photo-ATRP can be greatly improved by changing the chromophore and adding a copper cocatalyst. Using phenothiazine1-PTZ, a controlled radical polymerization reaction of methyl acrylate was achieved, with polydispersity as low as 1.04. These optimized results compare favorably to those using state-of-the-art conjugated microporous phenothiazine catalysts42, which are also significantly more cumbersome to process and to characterize in terms of active site structure and density.\nBy altering the excited-state redox potentials of catalysts, the preferred mechanistic pathway of a reaction can be switched. Previous studies of the hydroetherification of2dhave revealed that the use of different organic photocatalysts can engender divergent regiochemical outcomes32. Above, we showed that linear product is exclusively formed when using1-PDIas the photocatalyst. When, instead, the pyrene1-Pyrand long-wave UV (365 nm) irradiation were used, high branched selectivity was observed (3d’), attributable to the favorability of a photo-reductive, rather than oxidative, route.\nIn addition to the organic catalysts described above, organometallic dyes can also be easily incorporated through polymer post-functionalization. We were able to coordinate iridium to a porous bipyridine-based material (1-Bpy-Ir), producing chelated iridium complexes that resemble prototypical organometallic photocatalysts7. Iridium photocatalysts can initiate and maintain a Ni(I/III)-based catalytic cycle for Buchwald–Hartwig C–N cross-coupling43,44, a staple transformation in pharmaceutical research and development. Using a low loading of1-Bpy-Irin conjunction with catalytic NiBr2, the amination of an aryl bromide was accomplished in a clean and high-yielding manner (product4b), without the requirement for any additional ligand for nickel. Finally, as cationic photocatalysts are frequently employed to access the most oxidizing excited-state species, we aimed to demonstrate that charged chromophores such as acridiniums could be incorporated through post-polymerization alkylation. The polymer1-Acr-Mewas synthesized in high yield and found to be an extraordinarily active heterogeneous photocatalyst, promoting the direct acylation of aniline at nearly an order-of-magnitude reduced acridinium loading compared to previously reported homogenous reactions45.\nCollectively, our findings imply that a broad scope of photocatalytic chromophores, from common to yet-undiscovered, could be integrated by copolymerization into the rigid structural framework of1. As the scaffold consistently bestows high specific surface area, excellent film properties, and processability in select non-polar solvents, the described method may constitute a general strategy for the covalent heterogenization of catalysts into porous materials for efficient and economical organic synthesis. In particular, we anticipate that these porous-polymer films, in the role of photocatalytic stationary phases in continuous flow, could represent an enabling technology for large-scale chemical manufacturing.\n\nMethods\n\nA dry 20 mL scintillation vial, equipped with a magnetic stir bar, was charged sequentially with1-PDI(1.34 mg, 0.18 μmol, 0.018 mol% PDI), methanol (4.0 mL), and methyl phenyl sulfide (124.2 mg, 1.0 mmol, 1.0 equiv). The vial was submerged into an ultrasonic bath for 1 min to disperse the catalyst. The vial was capped with a screw cap containing a rubber septum, which was punctured with two 18-gauge needles to vent the reaction mixture to the atmosphere. The cloudy red suspension was stirred vigorously at rt under illumination from a 450 nm blue LED lamp for 6 h. At this point, both the irradiation and the stirring were stopped, and the catalyst allowed to settle for 30 min. Additional methanol (1.0 mL) was added, and the mixture was filtered through a short plug of Celite, washing with methanol (5.0 mL). The filtrate was concentrated with the aid of a rotary evaporator, and the title compound was obtained as a clear oil (133.1 mg, 95% yield) after column chromatography on silica gel, using DCM as the eluent. The catalytic polymer1-PDIcould be recovered from the Celite by washing with THF (10 mL).1H NMR(500 MHz, CDCl3) δ 7.67–7.61 (m, 2H), 7.55–7.47 (m, 3H), 2.72 (s, 3H).13C NMR(125 MHz, CDCl3) δ 145.9, 131.1, 129.5, 123.6, 44.1. The spectral data agreed closely with those reported in the literature46.\n\nData availability\n\nAdditional experimental methods and data are available in the Supplementary Information of this paper.\n\nReferences\n\nMcAtee, R. C., McClain, E. J. & Stephenson, C. R. J. Illuminating photoredox catalysis.Trends Chem.1, 111–125 (2019).\nCASArticleGoogle Scholar\nShaw, M. H., Twilton, J. & MacMillan, D. W. C. Photoredox catalysis in organic chemistry.J. Org. Chem.81, 6898–6926 (2016).\nCASPubMedPubMed CentralArticleGoogle Scholar\nRomero, N. A. & Nicewicz, D. A. Organic photoredox catalysis.Chem. Rev.116, 10075–10166 (2016).\nCASPubMedArticleGoogle Scholar\nSchultz, D. M. & Yoon, T. P. Solar synthesis: prospects in visible light photocatalysis.Science343, 1239176 (2014).\nPubMedPubMed CentralArticleCASGoogle Scholar\nPrier, C. K., Rankic, D. A. & MacMillan, D. W. C. Visible light photoredox catalysis with transition metal complexes: applications in organic synthesis.Chem. Rev.113, 5322–5363 (2013).\nCASPubMedPubMed CentralArticleGoogle Scholar\nFriedmann, D., Hakki, A., Kim, H., Choi, W. & Bahnemann, D. Heterogeneous photocatalytic organic synthesis: state-of-the-art and future perspectives.Green. Chem.18, 5391–5411 (2016).\nCASArticleGoogle Scholar\nPitre, S. P., Overman, L. E. Strategic use of visible-light photoredox catalysis in natural product synthesis.Chem. Rev.https://doi.org/10.1021/acs.chemrev.1c00247(2021).\nTwilton, J. et al. The merger of transition metal and photocatalysis.Nat. Rev. Chem.1, 0052 (2017).\nCASArticleGoogle Scholar\nLang, X., Zhao, J. & Chen, X. Cooperative photoredox catalysis.Chem. Soc. Rev.45, 3026–3038 (2016).\nCASPubMedArticleGoogle Scholar\nSkubi, K. L., Blum, T. R. & Yoon, T. P. Dual catalysis strategies in photochemical synthesis.Chem. Rev.116, 10035–10074 (2016).\nCASPubMedPubMed CentralArticleGoogle Scholar\nGisbertz, S. & Pieber, B. Heterogeneous photocatalysis in organic synthesis.ChemPhotoChem4, 456–475 (2020).\nCASArticleGoogle Scholar\nLe, C. et al. A general small-scale reactor to enable standardization and acceleration of photocatalytic reactions.ACS Cent. Sci.3, 647–653 (2017).\nCASPubMedArticleGoogle Scholar\nCambié, D., Bottecchia, C., Straathof, N. J. W., Hessel, V. & Noel, T. Application of continuous-flow photochemistry in organic synthesis, material science, and water treatment.Chem. Rev.116, 10276–10341 (2016).\nPubMedArticleCASGoogle Scholar\nHarper, K. C., Moschetta, E. G., Bordawekar, S. V. & Wittenberger, S. J. A laser driven flow chemistry platform for scaling photochemical reactions with visible light.ACS Cent. Sci.5, 109–115 (2019).\nCASPubMedPubMed CentralArticleGoogle Scholar\nBuglioni, L., Raymenants, F., Slattery, A., Zondag, S. D. A., Noël, T. Technological innovations in photochemistry for organic synthesis: flow chemistry, high-throughput experimentation, scale-up, and photoelectrochemistry.Chem. Rev.https://doi.org/10.1021/acs.chemrev.1c00332(2021).\nKisch, H. Semiconductor photocatalysis—mechanistic and synthetic aspects.Angew. Chem. Int. Ed.52, 812–847 (2013).\nCASArticleGoogle Scholar\nGhosh, I. et al. Organic semiconductor photocatalyst can bifunctionalize arenes and heteroarenes.Science365, 360–366 (2019).\nADSCASPubMedArticleGoogle Scholar\nZeng, L., Guo, X., He, C. & Duan, C. Metal−organic frameworks: versatile materials for heterogeneous photocatalysis.ACS Catal.6, 7935–7947 (2016).\nCASArticleGoogle Scholar\nWang, H. et al. Covalent organic framework photocatalysts: structures and applications.Chem. Soc. Rev.49, 4135–4165 (2020).\nCASPubMedArticleGoogle Scholar\nZhang, W. et al. Visible light-driven C-3 functionalization of indoles over conjugated microporous polymers.ACS Catal.8, 8084–8091 (2018).\nCASArticleGoogle Scholar\nMak, C. H. et al. Heterogenization of homogeneous photocatalysts utilizing synthetic and natural support materials.J. Mater. Chem. A9, 4454–4504 (2021).\nCASArticleGoogle Scholar\nGuo, S. & Swager, T. M. Versatile porous poly(arylene ether)s via Pd-catalyzed C–O polycondensation.J. Am. Chem. Soc.143, 11828–11835 (2021).\nCASPubMedArticleGoogle Scholar\nSwager, T. M. Iptycenes in the design of performance polymers.Acc. Chem. Res.41, 1181–1189 (2008).\nCASPubMedArticleGoogle Scholar\nLong, T. M. & Swager, T. M. Molecular design of free volume as a route to low-κ dielectric materials.J. Am. Chem. Soc.125, 14113–14119 (2003).\nCASPubMedArticleGoogle Scholar\nLong, T. M. & Swager, T. M. Using “internal free volume” to increase chromophore alignment.J. Am. Chem. Soc.127, 17976–17977 (2002).\nGoogle Scholar\nLessard, J. J. et al. Self-catalyzing photoredox polymerization for recyclable polymer catalysts.Polym. Chem.12, 2205–2209 (2021).\nCASArticleGoogle Scholar\nSmith, J. D. et al. Organopolymer with dual chromophores and fast charge-transfer properties for sustainable photocatalysis.Nat. Commun.10, 1837 (2019).\nADSPubMedPubMed CentralArticleCASGoogle Scholar\nRosso, C., Filippini, G. & Prato, M. Use of perylene diimides in synthetic photochemistry.Eur. J. Org. Chem.8, 1193–1200 (2021).\nArticleCASGoogle Scholar\nLiu, H., Shen, L., Cao, Z. & Li, X. Covalently linked perylenetetracarboxylic diimide dimers and trimers with rigid “J-type” aggregation structure.Phys. Chem. Chem. Phys.16, 16399–16406 (2014).\nCASPubMedArticleGoogle Scholar\nGhosh, I., Ghosh, T., Bardagi, J. I. & König, B. Reduction of aryl halides by consecutive visible light-induced electron transfer processes.Science346, 725–728 (2014).\nADSCASPubMedArticleGoogle Scholar\nGao, Y. et al. Visible-light photocatalytic aerobic oxidation of sulfides to sulfoxides with a perylene diimide photocatalyst.Org. Biomol. Chem.17, 7144–7149 (2019).\nCASPubMedArticleGoogle Scholar\nRosso, C., Filippini, G., Cozzi, P. G., Gualandi, A. & Prato, M. Highly performing iodoperfluoroalkylation of alkenes triggered by the photochemical activity of perylene diimides.ChemPhotoChem3, 193–197 (2019).\nCASArticleGoogle Scholar\nWeiser, M., Hermann, S., Penner, A. & Wagenknecht, H.-A. Photocatalytic nucleophilic addition of alcohols to styrenes in Markovnikov and anti-Markovnikov orientation.Beilstein J. Org. Chem.11, 568–575 (2015).\nCASPubMedPubMed CentralArticleGoogle Scholar\nCui, L. et al. Metal-Free direct C–H perfluoroalkylation of arenes and heteroarenes using a photoredox organocatalyst.Adv. Synth. Catal.355, 2203–2007 (2013).\nCASArticleGoogle Scholar\nMeyer, A. U., Slanina, T., Yao, C.-J. & König, B. Metal-free perfluoroarylation of visible light photoredox catalysis.ACS Catal.6, 369–375 (2016).\nCASArticleGoogle Scholar\nLee, J.-W. et al. Organotextile catalysis.Science341, 1225–1229 (2013).\nADSCASPubMedArticleGoogle Scholar\nBaumann, M., Moody, T. S., Smyth, M. & Wharry, S. A perspective on continuous flow chemistry in the pharmaceutical industry.Org. Process Res. Dev.24, 1802–1813 (2020).\nCASArticleGoogle Scholar\nAdamo, A. et al. On-demand continuous-flow production of pharmaceuticals in a compact, reconfigurable system.Science352, 61–67 (2016).\nADSCASPubMedArticleGoogle Scholar\nWilliams, J. D. & Kappe, C. O. Recent advances toward sustainable flow photochemistry.Curr. Op. Green. Sustain. Chem.25, 100351 (2020).\nArticleGoogle Scholar\nYang, A. et al. Heterogeneous photoredox flow chemistry for the scalable organosynthesis of fine chemicals.Nat. Commun.11, 1239 (2020).\nADSCASPubMedPubMed CentralArticleGoogle Scholar\nMiyake, G. M. & Thierot, J. C. Perylene as an organic photocatalyst for the radical polymerization of functionalized vinyl monomers through oxidative quenching with alkyl bromides and visible light.Macromolecules47, 8255–8261 (2014).\nADSCASArticleGoogle Scholar\nDadashi-Silab, S. et al. Conjugated cross-linked phenothiazines as green or red light heterogeneous photocatalysts for copper-catalyzed atom transfer radical polymerization.J. Am. Chem. Soc.143, 9630–9638 (2021).\nCASPubMedArticleGoogle Scholar\nTill, N. A., Tian, L., Dong, Z., Scholes, G. D. & MacMillan, D. W. C. Mechanistic analysis of metallaphotoredox C–N coupling: photocatalysis initiates and perpetuates Ni(I)/Ni(III) coupling activity.J. Am. Chem. Soc.142, 15830–15841 (2020).\nCASPubMedArticleGoogle Scholar\nCorcoran, E. B. et al. Aryl amination using ligand-free Ni(II) salts and photoredox catalysis.Science353, 279–283 (2016).\nADSCASPubMedPubMed CentralArticleGoogle Scholar\nSong, W., Dong, K. & Li, M. Visible light-induced amide bond formation.Org. Lett.22, 371–375 (2020).\nCASPubMedArticleGoogle Scholar\nDanahy, K. E., Styduhar, E. D., Fodness, A. M., Heckman, L. M. & Jamison, T. F. On-demand generation and use in continuous synthesis of the ambiphilic nitrogen source chloramine.Org. Lett.22, 8392–8395 (2020).\nCASPubMedArticleGoogle Scholar\nDownload references\n\nAcknowledgements\n\nThis research was supported by the National Science Foundation DMR-1809740 (T.M.S.) and the KAUST sensor project REP-2719 (T.M.S.). R.Y.L. is funded in part by US Army Combat Capabilities Development Command Soldier Center in Natick, MA. The textile symbol in Fig.1has been adapted with permission from Flaticon.com.\n\nAuthor information\n\nThese authors contributed equally: Richard Y. Liu, Sheng Guo.\nInstitute for Soldier Nanotechnologies, 500 Technology Square, Cambridge, MA, 02139, USA\nRichard Y. Liu, Sheng Guo, Shao-Xiong Lennon Luo & Timothy M. Swager\nDepartment of Chemistry, 77 Massachusetts Avenue, Cambridge, MA, 02139, USA\nRichard Y. Liu, Sheng Guo, Shao-Xiong Lennon Luo & Timothy M. Swager\nYou can also search for this author inPubMedGoogle Scholar\nYou can also search for this author inPubMedGoogle Scholar\nYou can also search for this author inPubMedGoogle Scholar\nYou can also search for this author inPubMedGoogle Scholar\nR.Y.L., S.G., and T.M.S. designed and planned the main experiments. S.G. synthesized and characterized the polymers. S.-X.L.L. carried out XPS studies and assisted with photoconductivity measurements. R.Y.L. performed photophysical characterization, catalytic studies, and other experiments. R.Y.L, S.G., and T.M.S. wrote the manuscript.\nCorrespondence toTimothy M. Swager.\n\nEthics declarations\n\nThe authors (R.Y.L., S.G., and T.M.S.) declare the following competing interests: a patent has been filed covering the materials and methods reported herein. All of the authors declare no other competing interests.\n\nPeer review\n\nNature Communicationsthanks the anonymous reviewer(s) for their contribution to the peer review of this work.\n\nAdditional information\n\nPublisher’s noteSpringer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.\n\nSupplementary information\n\n\nRights and permissions\n\nOpen AccessThis article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this license, visithttp://creativecommons.org/licenses/by/4.0/.\nReprints and Permissions\n\nAbout this article\n\nLiu, R.Y., Guo, S., Luo, SX.L.et al.Solution-processable microporous polymer platform for heterogenization of diverse photoredox catalysts.Nat Commun13,2775 (2022). https://doi.org/10.1038/s41467-022-29811-6\nDownload citation\nReceived:05 February 2022\nAccepted:23 March 2022\nPublished:27 May 2022\nDOI:https://doi.org/10.1038/s41467-022-29811-6\nAnyone you share the following link with will be able to read this content:\nSorry, a shareable link is not currently available for this article.\n\nProvided by the Springer Nature SharedIt content-sharing initiative\n\nComments\n\nBy submitting a comment you agree to abide by ourTermsandCommunity Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.\n\n",
    "keywords": [
        "λmax",
        "λem",
        "yield",
        "yet",
        "wondered"
    ],
    "url": "https://www.nature.com/articles/s41467-022-29811-6#Abs1"
}

[keyword density data]

{
    "article": "title\\n\\nSolution-processable microporous polymer platform for heterogenization of diverse photoredox catalysts\\nAbstract\\n\\nIn contemporary organic synthesis, substances that access strongly oxidizing and/or reducing states upon irradiation have been exploited to facilitate powerful and unprecedented transformations. However, the implementation of light-driven reactions in large-scale processes remains uncommon, limited by the lack of general technologies for the immobilization, separation, and reuse of these diverse catalysts. Here, we report a new class of photoactive organic polymers that combine the flexibility of small-molecule dyes with the operational advantages and recyclability of solid-phase catalysts. The solubility of these polymers in select non-polar organic solvents supports their facile processing into a wide range of heterogeneous modalities. The active sites, embedded within porous microstructures, display elevated reactivity, further enhanced by the mobility of excited states and charged species within the polymers. The independent tunability of the physical and photochemical properties of these materials affords a convenient, generalizable platform for the metamorphosis of modern photoredox catalysts into active heterogeneous equivalents.\\n\\nIntroduction\\n\\nPhotoredox catalysis refers to an energy-conversion process in which the absorption of photons by a substance triggers electron transfer. In the past decade, the application of this phenomenon to organic synthesis has evolved from an academic curiosity to an enabling technology widely adopted in pharmaceutical synthesis1,2,3,4,5,6,7. By coupling photoexcitation with other molecular processes, including cooperative catalytic schemes8,9,10, hundreds of powerful and previously inconceivable organic transformations have now been realized. However, compared to the rapid discovery of new photon-dependent reactions, the implementation of these processes on large scale has seen only limited progress due to the high cost of many photocatalysts, challenges in catalyst removal, and restricted light penetration in traditional batch reactors11. Thus, the conception of new technologies12,13,14,15that generally enhance photoredox catalysis on scale is a pressing goal that has attracted immense research effort.\\nPhotoredox catalysts can be divided into two major groups (Fig.1a): homogenous materials, which are soluble in the solvents typically used for photoredox catalysis, and heterogeneous, which operate as insoluble solids. Due to ease of synthesis and tunability, the vast majority of newly developed catalysts belong to the former group, which includes small-molecule organic dyes3, such as perylene diimides (PDIs), and organometallic complexes5, such as Ir(ppy)3. Meanwhile, heterogeneous catalysts have the advantages of facile removal from reaction mixtures and recyclability, which are important from both cost and sustainability perspectives6,11. Well-known examples include two-dimensional (2-D) surface photocatalysts, especially semiconductors16such as mesoporous graphitic carbon nitride (mpg-CN)17, and porous three-dimensional (3-D) constructs, such as metal-organic/covalent organic frameworks (MOFs/COFs)18,19,20. The appealing characteristics of these materials arrive at the expense of diversity and tunability: there is often no rational technique to finely alter their photochemical properties while retaining the bulk physical properties. Importantly, the inability of most heterogeneous catalysts to be cast into films or coatings from liquid or solution phase is a considerable practical limitation, particularly towards their synergistic application with emerging technologies such as continuous-flow synthesis10,11.\\naClassification of photoredox catalysts according to solubility and geometric dimension of active surface.bTransforming a broad scope of small-molecule photocatalysts into diverse heterogeneous materials through solution-processable polymers.\\nA strategy that allows synthetic chemists to simultaneously take advantage of the respective desirable properties of homogeneous and heterogeneous catalysts would be exceptionally empowering. We proposed that one solution might involve a general method for converting any member of the vast toolbox of soluble photocatalysts into equivalent insoluble materials, while maintaining the original photochemical attributes and activity. However, to our knowledge, no heterogenization approach exists with broad scope both in the kinds of dyes incorporated and in the types of catalytic materials accessible. Although several examples of photocatalyst immobilization on insoluble solids have been demonstrated21, each strategy addresses only a small subset of chromophore structures, and the vast majority of supported catalysts lack solution-processability, severely limiting the operational modes that are available (typically only particles or dispersions).\\nHerein, by extending our recently reported Pd-catalyzed polycondensation reaction22, we present a method that, in principle, allows for any organic or organometallic photocatalyst to be heterogenized through copolymerization with porosity-promoting bulky organic monomers (triptycenes and spirobifluorenes)23,24,25. We note that while 2D and 3D extended materials are well represented among heterogeneous photocatalysts, examples of one-dimensional (1D) photocatalysts, in which the active components form chains, are relatively rare26,27. Importantly, our linear-polymer structures, while insoluble in the solvents typically employed for photoredox catalysis, remain highly soluble in a few, select handling solvents (DCM and THF). As a result of this processability, our immobilization technique is not restricted to any particular target mode of heterogeneous catalysis but can conveniently access a wide range of known morphologies, including dispersions, films, coatings, textiles, and magnetic nanoparticles.\\nAside from solution-processability and broad generality, our 1D polyether platform offers additional valuable benefits compared to existing heterogeneous photoredox catalysts. First, the solubility of organic polymers enables simple characterization of structure, molecular weight distribution, photocatalyst loading, defect concentration, and batch-to-batch quality variations, all of which are often impractical to analyze for conventional insoluble catalysts. Second, traditional materials relying on non-covalent interactions for dye attachment, or labile covalent bonds such as Si–O or Al–O bonds, can exhibit poor chemical stability21. In contrast, the polymers developed in this work exhibit exceptional thermal stability and inertness toward a variety of chemical conditions. Third, the immobilization of small-molecule catalysts on heterogeneous supports often leads to drastically decreased turnover frequencies; however, with our polymer materials, due to a combination of reduced aggregation-induced quenching, highly porous-polymer structure, and energy- and charge-transport processes, very efficient catalysis can be achieved, in some cases with rates exceeding than with the corresponding monomers. Finally, the design of our polymers enables independent control of the physical and photochemical properties without alteration to the synthetic procedure.\\n\\nResults and discussion\\n\\nOur recent systematic study22showed that the foundational building blocks of our proposed material,t-Bu-triptycene hydroquinone1-Aand spirobifluorene dibromide1-B, could be cross-coupled to form polymer1with excellent molecular weight, polydispersity, porosity, and film properties (Fig.2a). Through slight alteration of these prior conditions, we were able to introduce chromophore-containing comonomers to generate materials with similar physical attributes as1, but with added photocatalytic capabilities. As a model, we selected the classic dye perylene diimide (PDI) as synthetically versatile redox-active moiety28. With the addition of 10% of PDI dibromide (1-C) in the polycondensation process, we obtained a near-quantitative yield of a deep-red solid (1-PDI) with high molecular weight and moderate polydispersity. The porosity was evaluated through the Brunauer–Emmett–Teller (BET) gas adsorption method, revealing a considerable specific surface area (SSA) of 338 m2/g. The essential role of the triptycene units in endowing the polymer with this exceptional porosity is evident from the very low SSA of related polymer1-PDI-HQ, in whichpara-phenylenes are present instead. Based on the N2adsorption isotherm, the estimated pore size distribution was concentrated in the microporous range (<2 nm, Supplementary Fig.3-1). Furthermore, the polymer showed excellent thermal and chemical stability under controlled conditions (Supplementary Fig.4-2). The extent of incorporation of the PDI was determined by1H NMR analysis to be 10%, in agreement with the fraction of monomer employed during the synthesis.\\naSynthesis of1,1-PDI, and1-PDI-HQ. See the Supplementary Information for detailed conditions.bPhotophysical properties.cLight-driven reversible redox process with an amine partner.dKinetics of catalytic photo-oxidation model reaction. BET = Brunauer–Emmett–Teller method.aToo low to be reliably determined from N2adsorption isotherm.bDilute THF solution (1.0 × 10–5–5.0 × 10–7M).cSolid state (powder).\\nThe absorption and steady-state photoluminescence spectra of1-PDIcontained distinctive features (λmax= 523 nm, λem= 528 nm in THF solution) characteristic of PDI units (Fig.2b). The solid-state photophysical properties of1-PDIshowed notable differences when compared against monomeric (PDI-Mono) and non-porous (1-PDI-HQ) analogues. Although the molecular dye displayed a very low fluorescence quantum yield in the solid state (QY = 0.02), the chromophores covalently hosted within triptycene-containing frameworks were bright emitters (QY = 0.29 for1-PDI), indicating a reduction in aggregation-based quenching. Monomeric PDIs are known to exhibit a substantial emission red-shift upon aggregation29, and we found that the difference in emission maxima between solution and solid states was substantially smaller for the polymers than for monomers and, in particular, smallest for the porous1-PDI.\\nAs a test of the photo-activated reactivity of these polymers, irradiation of a DCM solution of1-PDIwith excess Et3N resulted in isosbestic conversion to a new species, with spectral features indicative of PDI radical anions (Fig.2c)30. Exposure of the resulting solution to atmospheric oxygen regenerated the original polymer over the course of 2 h. Given its promising reversibly stoichiometric reactivity, a dispersion of the catalyst in methanol was evaluated in a simple aerobic photo-oxidation reaction31of methyl phenyl sulfide (2a, Fig.2d). At remarkably low loading of catalyst (0.018 mol%), full consumption of starting material could be observed after 6 h. Parallel experiments at equivalent optical densities of a non-porous analogue (1-PDI-HQ) and a monomeric PDI (PDI-Mono) proceeded at a significantly lower rate and only preceded to 80 and 76% conversion over the same time period. An additional control using a combination ofPDI-Monowith dye-free polyether1(premixed in dichloromethane solution, then dried and dispersed in methanol) showed a reaction rate similar to usingPDI-Monoalone. The reactions did not progress during a 30 min interval with no LED irradiation, confirming that the chosen transformation requires continuous photoirradiation. We hypothesized that, in addition to the steric insulation from aggregation-based quenching, the elevated activity of1-PDIis partially the result of the photoexcited state “hopping” between different dye centers, thereby enhancing the probability of encounter with a substrate or cocatalyst. Indeed, using fluorescence polarization spectroscopy, we observed quencher-dependent behavior consistent with considerable excited-state energy transfer between dye units (for extended discussion, see the Supplementary Information, Section 10.2, Figs.S7-2–S7-4). The mobility of charged states throughout the porous polymer can similarly contribute to an enhanced efficiency. In bulk conductivity measurements, we showed that irradiation of a film of1-PDIin the presence of triethylamine results in an instantaneous and reversible increase in conductivity (Supplementary Fig.7-5), suggesting facile transport of negative charges between PDI units.\\nTo illustrate the synthetic versatility of porous linear-polymer catalysts, we applied1-PDIto a representative range of modern photoredox transformations (Fig.3a). On a 1.0 mmol scale, the aerobic oxidation of thioether2ato sulfoxide3awas accomplished in 95% yield using an alcoholic dispersion of1-PDIand irradiation with a 450 nm LED. Photoreduction reactions could be performed through the in situ generation of PDI radical anion under inert atmosphere29. Using triethylamine as the terminal reductant, aryl bromide2bwas successfully hydrodehalogenated. The same catalyst could induce atom-transfer radical addition reactions to alkenes: fluorinated alkyl bromide3cwas prepared in high yield from 5-bromo-1-pentene (2c) and perfluorooctyl iodide32. Since3cis a non-commercial fluorosurfactant essential to other research in our laboratory, we pursued a scaled-up synthesis enabled by catalyst1-PDI. Over 12 g of product could be obtained in a single run, demonstrating the applicability of these porous materials in a preparative setting. Heterogeneous photoredox conditions were also applicable in polar addition to alkenes, as shown by the anti-Markovnikov hydroetherification of2d33. In another activation mode, the oxidative C–H functionalization of electron-rich arene2ewas conducted using potassium bromide as the bromine-atom source17. Both oxidative and reductive modes of radical generation were successful using1-PDIas a photocatalyst, and regioselective addition of these radicals to heteroarenes was accomplished (3fand3g)34,35.\\naPerformance of1-PDIacross various reaction classes.bRecyclable usage of1-PDIas a coating on a glass vial in the photo-oxidation of2a.cRecyclable usage of1-PDIdeposited on cotton in the photo-oxidation of2a.dCoated superparamagnetic silica nanoparticles.eContinuous-flow synthesis using a film of1-PDIin glass tubing.\\nMany valuable modes of heterogeneous catalysis are accessible due to the solubility profile of poly(arylene ether)s. Using a standard rotary evaporator, a thin film layer could be deposited on the bottom portion of a glass reaction vial, which has maximal exposure to the light source (Fig.3b). The coated vessel could be conveniently reused to perform the photo-oxidation of2a, with negligible change in reaction yield over five cycles. Textile-supported organocatalysts have recently been advanced as convenient promoters for acid- and base-dependent reactions36. We showed that1-PDIcan be robustly deposited onto cotton, and that the resulting fluorescent fabric is also a reusable photo-oxidation catalyst (Fig.3c). The polymer coating could adhere strongly to silica nanoparticles bearing a superparamagnetic iron oxide core (Fig.3d), which facilitates separation from reaction mixtures for which filtration is impractical. The resulting red powder was an effective agent in photoredox reactions, and the application of a handheld permanent magnet allowed for rapid and complete separation of the catalyst from the methanol solution (Supplementary Movie1). The catalytic layer could be instantly desorbed from the surface by immersion in tetrahydrofuran and thus recovered from the silica (Supplementary Movie2).\\nContinuous-flow synthesis, referring to processes performed within narrow tubing under constant material input and output, has attracted significant attention as an alternative to traditional batch chemical processes37,38. Flow systems are associated with superior heat exchange, mixing, reproducibility, and safety profiles39. The narrow-channel configuration of the reactors has proved particularly promising in large-scale photocatalysis applications, as enhanced light exposure is afforded with decreased risk of over-irradiation that can induce unwanted side-reactions40. A solution-processable material for the immobilization of diverse photocatalysts along the interior surface of the reactor tubing would provide two main advantages. First, no separation of the dyes from the mobile phase would be required after the reaction. Second, by concentrating the chromophores close to the exterior of the reaction mixture, light exposure could be enhanced. Taking advantage of the excellent film-forming characteristics of our porous polyethers, we showed that flow chemistry could be conducted using a photoredox-active coating as a stationary phase (Fig.3e). To the coolant coil of a spiral reflux condenser, a standard piece of equipment in laboratories that perform microscale organic synthesis, was applied a thin layer of1-PDIas a solution in THF, drying with gentle heating and air flow. With the system under constant irradiation with blue LEDs, reactants for the C–H bromination of2ewere pumped into the coil using a syringe pump (approximate residence time of 60 min). At steady state, a good yield of3ewas measured in the output stream.\\nA unique advantage of linear polymers as an immobilization platform is the ability to independently adjust the backbone-derived physical characteristics and the chromophore-derived photochemical attributes (Fig.4a). Achieving adhesion to fluoropolymer surfaces was desirable since inert, flexible substances such as perfluoroalkoxy (PFA) plastics are the most common tubing materials for modern continuous-flow synthesis. Since most poly(arylene ether) polymers such as1-PDIadsorb poorly to fluorinated surfaces, and therefore do not form high-quality films, we wondered if we could enhance the attractive interaction through post-polymerization functionalization of the backbone. We subjected the polymer to C–H fluoroalkylation conditions with photogenerated perfluoroalkyl (n-C17F35) radicals to obtain1-PDI-C17F35with relatively high fluorine content (F/C ratio = 1:4 by XPS). This new polymer was successfully deposited on narrow inner diameter (1/16 in) PFA tubing and effectively employed as a catalytic stationary phase in the flow synthesis of3e(Fig.4b).\\naConvenient and independent alteration of catalyst properties.bPost-polymerization functionalization of the backbone enables adhesion to fluorinated surfaces for photoredox catalysis.cStructural variation in the photocatalytic moiety: incorporation of common photoredox chromophores. Numbers in parentheses indicate photoluminescence quantum yields in dilute THF solution (1.0 × 10–5–5.0 × 10–7M).dExample reactions using modified photocatalysts.aMolecular weight distribution measured on1-Bpyprior to Ir incorporation. PMDETA = 1,1,4,7,7-pentamethyldiethylenetriamine.\\nThe identity of the covalently-included dye and its doping concentration could be adjusted arbitrarily with little effect on the porosity, solution-processability, and film-forming properties of the polymer. The substitution of other groups in lieu of PDI is synthetically straightforward and requires essentially no procedural modifications (Fig.4c). Perylenes are highly reducing photocatalysts that have been used in applications such as light-activated atom-transfer radical polymerization (photo-ATRP) reactions41. Using perylene-containing1-Peras a catalyst, the polymerization of methyl methacrylate could be performed using sunlight as the energy source. With minimal optimization, the poly(methyl methacrylate) (PMMA,4a) product could be obtained with good molecular weight and polydispersity. Furthermore, the catalytic polymer was easily separated from the product polymer, as the former is strictly insoluble in acetone. The outcome of photo-ATRP can be greatly improved by changing the chromophore and adding a copper cocatalyst. Using phenothiazine1-PTZ, a controlled radical polymerization reaction of methyl acrylate was achieved, with polydispersity as low as 1.04. These optimized results compare favorably to those using state-of-the-art conjugated microporous phenothiazine catalysts42, which are also significantly more cumbersome to process and to characterize in terms of active site structure and density.\\nBy altering the excited-state redox potentials of catalysts, the preferred mechanistic pathway of a reaction can be switched. Previous studies of the hydroetherification of2dhave revealed that the use of different organic photocatalysts can engender divergent regiochemical outcomes32. Above, we showed that linear product is exclusively formed when using1-PDIas the photocatalyst. When, instead, the pyrene1-Pyrand long-wave UV (365 nm) irradiation were used, high branched selectivity was observed (3d’), attributable to the favorability of a photo-reductive, rather than oxidative, route.\\nIn addition to the organic catalysts described above, organometallic dyes can also be easily incorporated through polymer post-functionalization. We were able to coordinate iridium to a porous bipyridine-based material (1-Bpy-Ir), producing chelated iridium complexes that resemble prototypical organometallic photocatalysts7. Iridium photocatalysts can initiate and maintain a Ni(I/III)-based catalytic cycle for Buchwald–Hartwig C–N cross-coupling43,44, a staple transformation in pharmaceutical research and development. Using a low loading of1-Bpy-Irin conjunction with catalytic NiBr2, the amination of an aryl bromide was accomplished in a clean and high-yielding manner (product4b), without the requirement for any additional ligand for nickel. Finally, as cationic photocatalysts are frequently employed to access the most oxidizing excited-state species, we aimed to demonstrate that charged chromophores such as acridiniums could be incorporated through post-polymerization alkylation. The polymer1-Acr-Mewas synthesized in high yield and found to be an extraordinarily active heterogeneous photocatalyst, promoting the direct acylation of aniline at nearly an order-of-magnitude reduced acridinium loading compared to previously reported homogenous reactions45.\\nCollectively, our findings imply that a broad scope of photocatalytic chromophores, from common to yet-undiscovered, could be integrated by copolymerization into the rigid structural framework of1. As the scaffold consistently bestows high specific surface area, excellent film properties, and processability in select non-polar solvents, the described method may constitute a general strategy for the covalent heterogenization of catalysts into porous materials for efficient and economical organic synthesis. In particular, we anticipate that these porous-polymer films, in the role of photocatalytic stationary phases in continuous flow, could represent an enabling technology for large-scale chemical manufacturing.\\n\\nMethods\\n\\nA dry 20 mL scintillation vial, equipped with a magnetic stir bar, was charged sequentially with1-PDI(1.34 mg, 0.18 μmol, 0.018 mol% PDI), methanol (4.0 mL), and methyl phenyl sulfide (124.2 mg, 1.0 mmol, 1.0 equiv). The vial was submerged into an ultrasonic bath for 1 min to disperse the catalyst. The vial was capped with a screw cap containing a rubber septum, which was punctured with two 18-gauge needles to vent the reaction mixture to the atmosphere. The cloudy red suspension was stirred vigorously at rt under illumination from a 450 nm blue LED lamp for 6 h. At this point, both the irradiation and the stirring were stopped, and the catalyst allowed to settle for 30 min. Additional methanol (1.0 mL) was added, and the mixture was filtered through a short plug of Celite, washing with methanol (5.0 mL). The filtrate was concentrated with the aid of a rotary evaporator, and the title compound was obtained as a clear oil (133.1 mg, 95% yield) after column chromatography on silica gel, using DCM as the eluent. The catalytic polymer1-PDIcould be recovered from the Celite by washing with THF (10 mL).1H NMR(500 MHz, CDCl3) δ 7.67–7.61 (m, 2H), 7.55–7.47 (m, 3H), 2.72 (s, 3H).13C NMR(125 MHz, CDCl3) δ 145.9, 131.1, 129.5, 123.6, 44.1. The spectral data agreed closely with those reported in the literature46.\\n\\nData availability\\n\\nAdditional experimental methods and data are available in the Supplementary Information of this paper.\\n\\nReferences\\n\\nMcAtee, R. C., McClain, E. J. & Stephenson, C. R. J. Illuminating photoredox catalysis.Trends Chem.1, 111–125 (2019).\\nCASArticleGoogle Scholar\\nShaw, M. H., Twilton, J. & MacMillan, D. W. C. Photoredox catalysis in organic chemistry.J. Org. Chem.81, 6898–6926 (2016).\\nCASPubMedPubMed CentralArticleGoogle Scholar\\nRomero, N. A. & Nicewicz, D. A. Organic photoredox catalysis.Chem. Rev.116, 10075–10166 (2016).\\nCASPubMedArticleGoogle Scholar\\nSchultz, D. M. & Yoon, T. P. Solar synthesis: prospects in visible light photocatalysis.Science343, 1239176 (2014).\\nPubMedPubMed CentralArticleCASGoogle Scholar\\nPrier, C. K., Rankic, D. A. & MacMillan, D. W. C. Visible light photoredox catalysis with transition metal complexes: applications in organic synthesis.Chem. Rev.113, 5322–5363 (2013).\\nCASPubMedPubMed CentralArticleGoogle Scholar\\nFriedmann, D., Hakki, A., Kim, H., Choi, W. & Bahnemann, D. Heterogeneous photocatalytic organic synthesis: state-of-the-art and future perspectives.Green. Chem.18, 5391–5411 (2016).\\nCASArticleGoogle Scholar\\nPitre, S. P., Overman, L. E. Strategic use of visible-light photoredox catalysis in natural product synthesis.Chem. Rev.https://doi.org/10.1021/acs.chemrev.1c00247(2021).\\nTwilton, J. et al. The merger of transition metal and photocatalysis.Nat. Rev. Chem.1, 0052 (2017).\\nCASArticleGoogle Scholar\\nLang, X., Zhao, J. & Chen, X. Cooperative photoredox catalysis.Chem. Soc. Rev.45, 3026–3038 (2016).\\nCASPubMedArticleGoogle Scholar\\nSkubi, K. L., Blum, T. R. & Yoon, T. P. Dual catalysis strategies in photochemical synthesis.Chem. Rev.116, 10035–10074 (2016).\\nCASPubMedPubMed CentralArticleGoogle Scholar\\nGisbertz, S. & Pieber, B. Heterogeneous photocatalysis in organic synthesis.ChemPhotoChem4, 456–475 (2020).\\nCASArticleGoogle Scholar\\nLe, C. et al. A general small-scale reactor to enable standardization and acceleration of photocatalytic reactions.ACS Cent. Sci.3, 647–653 (2017).\\nCASPubMedArticleGoogle Scholar\\nCambié, D., Bottecchia, C., Straathof, N. J. W., Hessel, V. & Noel, T. Application of continuous-flow photochemistry in organic synthesis, material science, and water treatment.Chem. Rev.116, 10276–10341 (2016).\\nPubMedArticleCASGoogle Scholar\\nHarper, K. C., Moschetta, E. G., Bordawekar, S. V. & Wittenberger, S. J. A laser driven flow chemistry platform for scaling photochemical reactions with visible light.ACS Cent. Sci.5, 109–115 (2019).\\nCASPubMedPubMed CentralArticleGoogle Scholar\\nBuglioni, L., Raymenants, F., Slattery, A., Zondag, S. D. A., Noël, T. Technological innovations in photochemistry for organic synthesis: flow chemistry, high-throughput experimentation, scale-up, and photoelectrochemistry.Chem. Rev.https://doi.org/10.1021/acs.chemrev.1c00332(2021).\\nKisch, H. Semiconductor photocatalysis—mechanistic and synthetic aspects.Angew. Chem. Int. Ed.52, 812–847 (2013).\\nCASArticleGoogle Scholar\\nGhosh, I. et al. Organic semiconductor photocatalyst can bifunctionalize arenes and heteroarenes.Science365, 360–366 (2019).\\nADSCASPubMedArticleGoogle Scholar\\nZeng, L., Guo, X., He, C. & Duan, C. Metal−organic frameworks: versatile materials for heterogeneous photocatalysis.ACS Catal.6, 7935–7947 (2016).\\nCASArticleGoogle Scholar\\nWang, H. et al. Covalent organic framework photocatalysts: structures and applications.Chem. Soc. Rev.49, 4135–4165 (2020).\\nCASPubMedArticleGoogle Scholar\\nZhang, W. et al. Visible light-driven C-3 functionalization of indoles over conjugated microporous polymers.ACS Catal.8, 8084–8091 (2018).\\nCASArticleGoogle Scholar\\nMak, C. H. et al. Heterogenization of homogeneous photocatalysts utilizing synthetic and natural support materials.J. Mater. Chem. A9, 4454–4504 (2021).\\nCASArticleGoogle Scholar\\nGuo, S. & Swager, T. M. Versatile porous poly(arylene ether)s via Pd-catalyzed C–O polycondensation.J. Am. Chem. Soc.143, 11828–11835 (2021).\\nCASPubMedArticleGoogle Scholar\\nSwager, T. M. Iptycenes in the design of performance polymers.Acc. Chem. Res.41, 1181–1189 (2008).\\nCASPubMedArticleGoogle Scholar\\nLong, T. M. & Swager, T. M. Molecular design of free volume as a route to low-κ dielectric materials.J. Am. Chem. Soc.125, 14113–14119 (2003).\\nCASPubMedArticleGoogle Scholar\\nLong, T. M. & Swager, T. M. Using “internal free volume” to increase chromophore alignment.J. Am. Chem. Soc.127, 17976–17977 (2002).\\nGoogle Scholar\\nLessard, J. J. et al. Self-catalyzing photoredox polymerization for recyclable polymer catalysts.Polym. Chem.12, 2205–2209 (2021).\\nCASArticleGoogle Scholar\\nSmith, J. D. et al. Organopolymer with dual chromophores and fast charge-transfer properties for sustainable photocatalysis.Nat. Commun.10, 1837 (2019).\\nADSPubMedPubMed CentralArticleCASGoogle Scholar\\nRosso, C., Filippini, G. & Prato, M. Use of perylene diimides in synthetic photochemistry.Eur. J. Org. Chem.8, 1193–1200 (2021).\\nArticleCASGoogle Scholar\\nLiu, H., Shen, L., Cao, Z. & Li, X. Covalently linked perylenetetracarboxylic diimide dimers and trimers with rigid “J-type” aggregation structure.Phys. Chem. Chem. Phys.16, 16399–16406 (2014).\\nCASPubMedArticleGoogle Scholar\\nGhosh, I., Ghosh, T., Bardagi, J. I. & König, B. Reduction of aryl halides by consecutive visible light-induced electron transfer processes.Science346, 725–728 (2014).\\nADSCASPubMedArticleGoogle Scholar\\nGao, Y. et al. Visible-light photocatalytic aerobic oxidation of sulfides to sulfoxides with a perylene diimide photocatalyst.Org. Biomol. Chem.17, 7144–7149 (2019).\\nCASPubMedArticleGoogle Scholar\\nRosso, C., Filippini, G., Cozzi, P. G., Gualandi, A. & Prato, M. Highly performing iodoperfluoroalkylation of alkenes triggered by the photochemical activity of perylene diimides.ChemPhotoChem3, 193–197 (2019).\\nCASArticleGoogle Scholar\\nWeiser, M., Hermann, S., Penner, A. & Wagenknecht, H.-A. Photocatalytic nucleophilic addition of alcohols to styrenes in Markovnikov and anti-Markovnikov orientation.Beilstein J. Org. Chem.11, 568–575 (2015).\\nCASPubMedPubMed CentralArticleGoogle Scholar\\nCui, L. et al. Metal-Free direct C–H perfluoroalkylation of arenes and heteroarenes using a photoredox organocatalyst.Adv. Synth. Catal.355, 2203–2007 (2013).\\nCASArticleGoogle Scholar\\nMeyer, A. U., Slanina, T., Yao, C.-J. & König, B. Metal-free perfluoroarylation of visible light photoredox catalysis.ACS Catal.6, 369–375 (2016).\\nCASArticleGoogle Scholar\\nLee, J.-W. et al. Organotextile catalysis.Science341, 1225–1229 (2013).\\nADSCASPubMedArticleGoogle Scholar\\nBaumann, M., Moody, T. S., Smyth, M. & Wharry, S. A perspective on continuous flow chemistry in the pharmaceutical industry.Org. Process Res. Dev.24, 1802–1813 (2020).\\nCASArticleGoogle Scholar\\nAdamo, A. et al. On-demand continuous-flow production of pharmaceuticals in a compact, reconfigurable system.Science352, 61–67 (2016).\\nADSCASPubMedArticleGoogle Scholar\\nWilliams, J. D. & Kappe, C. O. Recent advances toward sustainable flow photochemistry.Curr. Op. Green. Sustain. Chem.25, 100351 (2020).\\nArticleGoogle Scholar\\nYang, A. et al. Heterogeneous photoredox flow chemistry for the scalable organosynthesis of fine chemicals.Nat. Commun.11, 1239 (2020).\\nADSCASPubMedPubMed CentralArticleGoogle Scholar\\nMiyake, G. M. & Thierot, J. C. Perylene as an organic photocatalyst for the radical polymerization of functionalized vinyl monomers through oxidative quenching with alkyl bromides and visible light.Macromolecules47, 8255–8261 (2014).\\nADSCASArticleGoogle Scholar\\nDadashi-Silab, S. et al. Conjugated cross-linked phenothiazines as green or red light heterogeneous photocatalysts for copper-catalyzed atom transfer radical polymerization.J. Am. Chem. Soc.143, 9630–9638 (2021).\\nCASPubMedArticleGoogle Scholar\\nTill, N. A., Tian, L., Dong, Z., Scholes, G. D. & MacMillan, D. W. C. Mechanistic analysis of metallaphotoredox C–N coupling: photocatalysis initiates and perpetuates Ni(I)/Ni(III) coupling activity.J. Am. Chem. Soc.142, 15830–15841 (2020).\\nCASPubMedArticleGoogle Scholar\\nCorcoran, E. B. et al. Aryl amination using ligand-free Ni(II) salts and photoredox catalysis.Science353, 279–283 (2016).\\nADSCASPubMedPubMed CentralArticleGoogle Scholar\\nSong, W., Dong, K. & Li, M. Visible light-induced amide bond formation.Org. Lett.22, 371–375 (2020).\\nCASPubMedArticleGoogle Scholar\\nDanahy, K. E., Styduhar, E. D., Fodness, A. M., Heckman, L. M. & Jamison, T. F. On-demand generation and use in continuous synthesis of the ambiphilic nitrogen source chloramine.Org. Lett.22, 8392–8395 (2020).\\nCASPubMedArticleGoogle Scholar\\nDownload references\\n\\nAcknowledgements\\n\\nThis research was supported by the National Science Foundation DMR-1809740 (T.M.S.) and the KAUST sensor project REP-2719 (T.M.S.). R.Y.L. is funded in part by US Army Combat Capabilities Development Command Soldier Center in Natick, MA. The textile symbol in Fig.1has been adapted with permission from Flaticon.com.\\n\\nAuthor information\\n\\nThese authors contributed equally: Richard Y. Liu, Sheng Guo.\\nInstitute for Soldier Nanotechnologies, 500 Technology Square, Cambridge, MA, 02139, USA\\nRichard Y. Liu, Sheng Guo, Shao-Xiong Lennon Luo & Timothy M. Swager\\nDepartment of Chemistry, 77 Massachusetts Avenue, Cambridge, MA, 02139, USA\\nRichard Y. Liu, Sheng Guo, Shao-Xiong Lennon Luo & Timothy M. Swager\\nYou can also search for this author inPubMedGoogle Scholar\\nYou can also search for this author inPubMedGoogle Scholar\\nYou can also search for this author inPubMedGoogle Scholar\\nYou can also search for this author inPubMedGoogle Scholar\\nR.Y.L., S.G., and T.M.S. designed and planned the main experiments. S.G. synthesized and characterized the polymers. S.-X.L.L. carried out XPS studies and assisted with photoconductivity measurements. R.Y.L. performed photophysical characterization, catalytic studies, and other experiments. R.Y.L, S.G., and T.M.S. wrote the manuscript.\\nCorrespondence toTimothy M. Swager.\\n\\nEthics declarations\\n\\nThe authors (R.Y.L., S.G., and T.M.S.) declare the following competing interests: a patent has been filed covering the materials and methods reported herein. All of the authors declare no other competing interests.\\n\\nPeer review\\n\\nNature Communicationsthanks the anonymous reviewer(s) for their contribution to the peer review of this work.\\n\\nAdditional information\\n\\nPublisher’s noteSpringer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.\\n\\nSupplementary information\\n\\n\\nRights and permissions\\n\\nOpen AccessThis article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this license, visithttp://creativecommons.org/licenses/by/4.0/.\\nReprints and Permissions\\n\\nAbout this article\\n\\nLiu, R.Y., Guo, S., Luo, SX.L.et al.Solution-processable microporous polymer platform for heterogenization of diverse photoredox catalysts.Nat Commun13,2775 (2022). https://doi.org/10.1038/s41467-022-29811-6\\nDownload citation\\nReceived:05 February 2022\\nAccepted:23 March 2022\\nPublished:27 May 2022\\nDOI:https://doi.org/10.1038/s41467-022-29811-6\\nAnyone you share the following link with will be able to read this content:\\nSorry, a shareable link is not currently available for this article.\\n\\nProvided by the Springer Nature SharedIt content-sharing initiative\\n\\nComments\\n\\nBy submitting a comment you agree to abide by ourTermsandCommunity Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.\\n\\n",
    "date": "2022-06-17",
    "keywords": {
        "yield": [
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            1,
            0,
            0,
            1,
            0,
            0,
            0,
            0,
            0,
            1,
            1,
            0,
            0,
            1,
            0,
            0,
            0,
            0,
            1,
            0,
            0,
            1,
            0,
            0,
            0,
            0,
            0,
            1,
            1,
            0,
            0,
            1,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0
        ],
        "λem": [
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            1,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0
        ],
        "λmax": [
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            1,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0,
            0
        ]
    },
    "url": "https://www.nature.com/articles/s41467-022-29811-6"
}

변경사항

[AI]

  • 기존의 Abstract만을 가지고 진행한 부분을 결과 발표 이후, 여러번의 실험결과, 가장 좋은 결과가 나온 reference 등을 제외한 본문만을 기반으로 분석을 진행하였습니다

[기술 스텍]

  • react → flutter(react대신 flutter web을 기반으로 작성되었습니다.)

[로그인, 회원가입]

  • 개발 이슈로 인해 해당 부분을 향후 개발 사항으로 미루고, 원래 있던 버튼은 github routing 버튼으로 변경하였습니다.

[미니맵]

  • flutter로 기술 스택이 변경되며, front상 미니맵 구현에 많은 에러 사항을 겪어 후속개발로 진행될 예정입니다.

향후 개발 진행사항

  • fast api 배포
  • 로그인, 회원가입 배포
  • minimap 기능 및 density graph 논문과의 배치 정확도 개선
  • nature communication 이외 데이터 대은
  • pdf 데이터 대응

unix_pongpong's People

Contributors

skyriver228 avatar izen1231 avatar afklee avatar

Watchers

 avatar  avatar

Recommend Projects

  • React photo React

    A declarative, efficient, and flexible JavaScript library for building user interfaces.

  • Vue.js photo Vue.js

    🖖 Vue.js is a progressive, incrementally-adoptable JavaScript framework for building UI on the web.

  • Typescript photo Typescript

    TypeScript is a superset of JavaScript that compiles to clean JavaScript output.

  • TensorFlow photo TensorFlow

    An Open Source Machine Learning Framework for Everyone

  • Django photo Django

    The Web framework for perfectionists with deadlines.

  • D3 photo D3

    Bring data to life with SVG, Canvas and HTML. 📊📈🎉

Recommend Topics

  • javascript

    JavaScript (JS) is a lightweight interpreted programming language with first-class functions.

  • web

    Some thing interesting about web. New door for the world.

  • server

    A server is a program made to process requests and deliver data to clients.

  • Machine learning

    Machine learning is a way of modeling and interpreting data that allows a piece of software to respond intelligently.

  • Game

    Some thing interesting about game, make everyone happy.

Recommend Org

  • Facebook photo Facebook

    We are working to build community through open source technology. NB: members must have two-factor auth.

  • Microsoft photo Microsoft

    Open source projects and samples from Microsoft.

  • Google photo Google

    Google ❤️ Open Source for everyone.

  • D3 photo D3

    Data-Driven Documents codes.