首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 718 毫秒
1.
Chiral N‐(binaphthyl‐2‐yl)thiophosphoramide L7 [O,O‐diethyl 2′‐(ethylamino)‐1,1′‐binaphthyl‐2‐ylamidothiophosphate] prepared from the reaction of diethyl chlorothiophosphate with (R)‐(+)‐N‐ethyl‐1,1′‐binaphthyl‐2,2′‐diamine was used as a catalytic chiral ligand in the first Cu(OTf)2‐promoted catalytic asymmetric addition of diethylzinc to N‐(diphenylphosphinoyl) imines in which ~85% ee can be realized.  相似文献   

2.
Several MOP ligands 5 containing aryl groups at 2′ position of (R)‐2‐(diphenylphosphino)‐1,1′‐binaphthyl skeleton were prepared and used for palladium‐catalyzed asymmetric hydrosilylation of cyclic 1,3‐dienes 6 with trichlorosilane. Highest enantioselectivity was observed in the reaction of 1,3‐cyclopentadiene ( 6a ) catalyzed by a palladium complex (0.25 mol %) coordinated with (R)‐2‐(diphenylphosphino)‐2′‐(3,5‐dimethyl‐4‐methoxyphenyl)‐1,1′‐binaphthyl ( 5f ), which gave (S)‐3‐(trichlorosilyl)cyclopentene of 90% ee.  相似文献   

3.
A highly efficient and enantioselective hydrogenation of unprotected β‐ketoenamines catalyzed with ruthenium(II) dichloro{(S)‐(−)‐2,2′‐bis[di(3,5‐xylyl)phosphino]‐1,1′‐binaphthyl}[(2S)‐(+)‐1,1‐bis(4‐methoxyphenyl)‐3‐methyl‐1,2‐butanediamine] {Ru[(S)‐xylbinap][(S)‐daipen]Cl2} has been successfully developed. This methodology provides a straightforward access to free γ‐secondary amino alcohols, which are key building blocks for a variety of pharmaceuticals and natural products, with high yields (>99%) and excellent enantioselectivities (up to 99% ee) in all cases.  相似文献   

4.
To develop conjugated polymers with low bandgap, deep HOMO level, and good solubility, a new conjugated alternating copolymer PC‐DODTBT based on N‐9′‐heptadecanyl‐2,7‐carbazole and 5, 6‐bis(octyloxy)‐4,7‐di(thiophen‐2‐yl)benzothiadiazole was synthesized by Suzuki cross‐coupling polymerization reaction. The polymer reveals excellent solubility and thermal stability with the decomposition temperature (5% weight loss) of 327°C. The HOMO level of PC‐DODTBT is ‐5.11 eV, indicating that the polymer has relatively deep HOMO level. The hole mobility of PC‐DODTBT as deduced from SCLC method was found to be 2.03 × 10?4 cm2/Versus Polymer solar cells (PSCs) based on the blends of PC‐DODTBT and [6,6]‐phenyl‐C71‐butyric acid methyl ester (PC71BM) with a weight ratio of 1:2.5 were fabricated. Under AM 1.5 (AM, air mass), 100 mW/cm?2 illumination, the devices were found to exhibit an open‐circuit voltage (Voc) of 0.73 V, short‐circuit current density (Jsc) of 5.63 mA/cm?2, and a power conversion efficiency (PCE) of 1.44%. This photovoltaic performance indicates that the copolymer is promising for polymer solar cells applications. © 2011 Wiley Periodicals, Inc. J Appl Polym Sci, 2011  相似文献   

5.
New aromatic diimide‐dicarboxylic acids having kinked and cranked structures, 2,2′‐bis(4‐trimellitimidophenoxy)biphenyl (2a) and 2,2′‐bis(4‐trimellitimidophenoxy)‐1,1′‐binaphthyl (2b), were synthesized by the reaction of trimellitic anhydride with 2,2′‐bis(4‐aminophenoxy)biphenyl (1a) and 2,2′‐bis(4‐aminophenoxy)‐1,1′‐binaphthyl (1b), respectively. Compounds 2a and 2b were characterized by FT‐IR and NMR spectroscopy and elemental analyses. Then, a series of novel aromatic poly(amide‐imide)s were prepared by the phosphorylation polycondensation of the synthesized monomers with various aromatic diamines. Owing to structural similarity, and a comparison of the characterization data, a model compound was synthesized by the reaction of 2b with aniline. The resulting polymers with inherent viscosities of 0.58–0.97 dl g?1 were obtained in high yield. The polymers were fully characterized by FT‐IR and NMR spectroscopy. The ultraviolet λmax values of the poly(amide‐imide)s were also determined. The polymers were readily soluble in polar aprotic solvents. They exhibited excellent thermal stabilities and had 10% weight loss at temperatures above 500 °C under a nitrogen atmosphere. Copyright © 2003 Society of Chemical Industry  相似文献   

6.
(11bR,11′bR)‐4,4′‐(1,2‐Phenylene)bis[4,5‐dihydro‐3H‐dinaphtho[2,1‐c:1′,2′‐e]phosphepin] [abbreviated as (R)‐BINAPHANE], (3R,3′R,4S,4′S,11bS,11′bS)‐4,4′‐bis(1,1‐dimethylethyl)‐4,4′,5,5′‐tetrahydro‐3,3′‐bi‐3H‐dinaphtho[2,1‐c:1′,2′‐e]phosphepin [(S)‐BINAPINE], (1S,1′S,2R,2′R)‐1,1′‐bis(1,1‐dimethylethyl)‐2,2′‐biphospholane [(S,S,R,R)‐TANGPHOS] and (2R,2′R,5R,5′R)‐1,1′‐(1,2‐phenylene)bis[2,5‐bis(1‐methylethyl)phospholane] [(R,R)‐i‐Pr‐DUPHOS] are C2‐bridged chiral diphosphines that form stable complexes with palladium(II) and platinum(II) containing a five‐membered chelate ring. The Pd(II)‐BINAPHANE catalyst displayed good to excellent enantioselectivities with ee values as high as 99.0% albeit in low yields for the carbonyl‐ene reaction between phenylglyoxal and alkenes. Its Pt(II) counterpart afforded improved yields while retaining satisfactory enantioselectivity. For the carbonyl‐ene reaction between ethyl trifluoropyruvate and alkenes, the Pd(II)‐BINAPHANE catalyst afforded both good yields and extremely high enantioselectivities with ees as high as 99.6%. A comparative study on the Pd(II) catalysts of the four C2‐bridged chiral diphosphines revealed that Pd(II)‐BINAPHANE afforded the best enantioselectivity. The ee values derived from Pd(II)‐BINAPHANE are much higher than those derived from the other three Pd(II) catalysts. A comparison of the catalyst structures shows that the Pd(II)‐BINAPHANE catalyst is the only one that has two bulky (R)‐binaphthyl groups close to the reaction site. Hence it creates a deep chiral space that can efficiently control the reaction behavior in the carbonyl‐ene reactions resulting in excellent enantioselectivity.  相似文献   

7.
(S)‐3‐Vinyl‐2,2′‐bisalkoxy‐1,1′‐binaphthyl ( 3 ) was synthesized via the Wittig reaction. Radical polymerization of all the monomers can take place smoothly in the temperature region tested. These polymers (poly‐ 3 ) showed very large specific optical rotations which were four times as large as those of the corresponding monomers 3 . Poly‐ 3 displayed optical rotations and Cotton effects in the UV?visible absorption region of side groups which were different from the corresponding monomers 3 and the model compounds (S)‐3‐ethyl‐2,2′‐bisalkoxy‐1,1′‐binaphthyl ( 4 ). These facts imply the formation of helicity of the main chain and the helical conformations were quite stable demonstrated by the unchanged pattern of temperature variable circular dichroism spectra. © 2015 Society of Chemical Industry  相似文献   

8.
A novel chiral H8‐1,1′‐binaphthyl‐based amino alcohol ligand (1Ra,2S,3R)‐ 2 has been synthesized and applied in the direct nucleophilic addition of organozincs (alkynylzinc and arylzinc prepared in situ) to aldehydes, yielding the corresponding optically active propargylic alcohols and diarylmethanols in high yields and good to excellent enantioselectivities. For the asymmetric arylation reaction, one catalyst (1Ra,2S,3R)‐ 2 can afford both enantiomers of many pharmaceutically interesting diarylmethanols by a proper combination of various arylzinc reagents and aldehydes.  相似文献   

9.
A new asymmetric catalytic system for the Mannich reaction of aldimines with trimethylsiloxyfuran is described. The combination of an axially chiral phosphine‐oxazoline ligand (S)‐2‐[(R)‐2′‐(diphenylphosphino)‐1,1′‐binaphthyl‐2‐yl]‐4‐phenyl‐4,5‐dihydrooxazole with silver acetate and 2,2,2‐trifluoroacetic acid is a very effective catalytic system in the asymmetric Mannich reaction of various aldimines with trimethylsiloxyfuran in dichloromethane at −78 °C, affording the corresponding adducts in up to 99% yield, 99:1 (dr) and 99% ee (major diastereoisomer) under mild conditions.  相似文献   

10.
A continuous‐flow process for the asymmetric hydrogenation of methyl propionylacetate as a prototypical β‐keto ester in a biphasic system of ionic liquid and supercritical carbon dioxide (scCO2) is presented. An established ruthenium/2,2′‐bis(diphenylphosphino)‐1,1′‐binaphthyl (BINAP) catalyst was immobilised in an imidazolium‐based ionic liquid while scCO2 was used as mobile phase transporting reactants in and products out of the reactor. The use of acidic additives led to significantly higher reaction rates and enhanced catalyst stability albeit at slightly reduced enantioselectivity. High single pass conversions (>90%) and good enantioselectivity (80–82% ee) were achieved in the first 80 h. The initial catalyst activity was retained to 91% after 100 h and to 69% after 150 h time‐on‐stream, whereas the enantioselectivity remained practically constant during the entire process. A total turnover number of ∼21,000 and an averaged space‐time yield (STYav) of 149 g L−1 h−1 were reached in a long‐term experiment. No ruthenium and phosphorus contaminants could be detected via inductively coupled plasma optical emission spectrometry (ICP‐OES) in the product stream and almost quantitative retention by the analysis of the stationary phase was confirmed. A comparison between batch‐wise and continuous‐flow operation on the basis of these data is provided.  相似文献   

11.
We have developed an efficient copper‐catalyzed method for the synthesis of N‐alkylanthranilic acids. The protocol uses inexpensive copper(I) iodide/racemic 1,1′‐binaphthyl‐2,2′‐diol (rac‐BINOL) as the catalyst/ligand system, readily available 2‐halobenzoic acids and aliphatic amines as the starting materials, the coupling reactions were performed at room temperature, and various functionalities in the substrates were tolerated.  相似文献   

12.
Novel phosphorus‐containing binaphthyl epoxy DGEBN (diglycidyl ether of 2,2′‐hydroxy‐1,1′‐binaphthalene) with high thermal performance was obtained from the addition reaction of DGEBN and diethyl phosphite. The modified binaphthyl epoxy was characterized by Fourier transform infrared (FTIR) and nuclear magnetic resonance spectroscopy. The dynamic mechanical property of the cured epoxy polymer was investigated by dynamic mechanical thermal analysis. The result revealed that the cured polymer with lower phosphorus content displayed higher value of the storage modulus when the networks reached rubbery state (above the glass transition temperature Tg). The Tgs decreased slightly with increasing phosphorous content. The thermal degradation was studied with thermogravimetric analysis and the evolved gas was analyzed using thermogravimetric analysis/Fourier transform infrared technique. The influence of phosphorus content and the chemical structure on the degradation behavior was discussed. The P‐modified binaphthyl epoxy polymers exhibited higher thermal stability than the P‐modified diglycidyl ether of bisphenol A polymer. Flammability measurements were performed by the examination of limited oxygen index and UL‐94 test. Compared with unmodified DGEBN, P‐containing epoxy polymers displayed higher limited oxygen index values and exhibited better flame retardance. POLYM. ENG. SCI., 2009. © 2009 Society of Plastics Engineers  相似文献   

13.
A series of new recyclable chiral dicationic chromium(III) salen complexes 1 – 10 bearing different substituents, viz., hydrogen, methyl, tert‐butyl, triphenylphosphinomethyl, triethylaminomethyl, methylimidazolium, methylpyridinium, methyl‐N,N‐dimethylpyridinium at the 3,3′‐ and 5,5′‐ positions of the salen unit with (1S,2S)(+)‐1,2‐diaminocyclohexane, (1S,2S)‐(−)‐1,2‐diphenyl‐1,2‐diaminoethane, and (S)(−)‐1,1′‐binaphthyl‐2,2′‐diamine collars have been synthesized and characterized by various physico‐chemical methods. These complexes were used as catalysts for the highly enantioselective aminolytic kinetic resolution of racemic trans‐epoxides with different anilines as nucleophiles at room temperature. With the use of catalyst 3 , anti‐β‐amino alcohols were obtained in excellent yields (>99% with respect to the nucleophile) and enantioselectivities (ee>99%) with the concomitant recovery of corresponding epoxides in high optical purity (ee up to >99%) and quantitative yields in 12 h. The catalyst 3 is recyclable in the aminolytic kinetic resolution process and worked well up to six cycles with retention of enantioselectivity.  相似文献   

14.
Conjugated polymers are highly desirable for the photovoltaic applications. We report the synthesis, characterization, optoelectronic properties, and solar cell application of two polymers, namely, poly[(9,9‐didodecylfluorene‐2,7‐diyl)‐alt‐(2,2′:5′,2″‐terthiophene‐5,5″‐diyl)] (P1) and poly[(1,4‐bis(dodecyloxy)benzene‐2,5‐diyl)‐alt‐(2,2′:5′,2″‐terthiophene‐5,5″‐diyl)] (P2). The polymers were synthesized via Stille cross‐coupling reaction, and were characterized by the gel permeation chromatography, nuclear magnetic resonance, Fourier transform infrared, UV–vis, thermogravimetric analysis, and cyclic voltammetry analyses. The two copolymers are processable due to their good solubility in organic solvents (tetrahydrofuran, CHCl3, toluene, chlorobenzene, and o‐dichlorobenzene). The optical band gaps (UV–vis, film, and Egopt) of the P1 and P2 are 2.04 and 2.00 eV, respectively. The density functional theory output structures showed that S O space interaction is likely responsible for the higher planarity of P2. The polymers showed low HOMO energy levels (P1: −5.33 eV, P2: −5.05 eV). The EHOMO for P1 is close to the EHOMO (−5.4 eV) of an ideal polymer, which is an important, rare, and main origin of the observed higher Voc (801–808 mV). The onset decomposition temperatures (Td) for the P1 and P2 are 418°C and 365°C, respectively. The polymer solar cell based on the P1: C60 (1: 1) and P2: C60 (1: 1) blend showed a power conversion efficiency (PCE) of 0.94 and 0.71%, respectively. The composite polymer : PC60BM = 1 : 2 increased PCE of the P1 (1.65%) and P2 (1.09%) under AM 1.5 illumination (100 mW/cm2). The study provided important examples to design donor–donor (D–D) polymers for the photovoltaic applications. © 2015 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2015 , 132, 42147.  相似文献   

15.
Palladium‐catalysed monophosphorylation of (R)‐2,2′‐bisperfluoroalkanesulfonates of BINOL (RF=CF3 or C4F9) by a diaryl phosphinate [Ar2P(O)H] followed by phosphine oxide reduction (Cl3SiH) then lithium diisopropylamide‐mediated anionic thia‐Fries rearrangement furnishes enantiomerically‐pure (R)‐2′‐diarylphosphino‐2′‐hydroxy‐3′‐perfluoralkanesulfonyl‐1,1′‐binaphthalenes [(R)‐ 8ab and (R)‐ 8g–j ], which can be further diversified by Grignard reagent (RMgX)‐mediated CF3‐displacement [→(R)‐ 8c–f ]. Coupling of (R)‐ 8a–j with (S)‐1,1′‐binaphthalene‐2,2′‐dioxychlorophosphine (S)‐ 9 generates 3′‐sulfonyl BINAPHOS ligands (R,S)‐ 10a–j in good yields (43–82%). These new ligands are of utlility in the asymmetric hydrophosphonylation of styrene ( 1 ) by 4,4,5,5‐tetramethyl‐1,3,2‐dioxaphospholane 2‐oxide ( 2 ), for which a combination of the chiral ligands with either [Pd(Cp)(allyl)] or [Pd(allyl)(MeCN)2]+/NaCH(CO2Me)2 proves to be a convenient and active pre‐catalyst system. A combination of an electron‐rich phosphine moiety and an electron‐deficient 3′‐sulfone moiety provides the best enantioselectivity to date for this process, affording the branched 2‐phenethenephosphonate, (−)‐iso‐ 3 , in up to 74% ee with ligand (R,S)‐ 10i , where Ar=p‐anisyl and the 3′‐SO2R group is triflone.  相似文献   

16.
Embedding metallic nanoparticles (MNPs) in organic solar cells (OSCs) is proposed as one of the promising strategies to enhance their photovoltaic performance owing to localized surface plasmon resonance, light scattering effects or a synergy of both effects derived from the MNPs. However, it has been demonstrated that MNPs wrapped by a thin dielectric silica shell can lead to better photovoltaic yield than bare MNPs due to the presence of the dielectric shell which avoids direct contact between the active layer and the MNPs, reducing the charge recombination and the exciton quenching loss at the metal surface. In this study, we report an alternative solution using an ultrathin dielectric layer coating silver nanoparticles (Ag NPs) for improving the performance of plasmonic inverted OSCs instead of the use of metal–dielectric core–shell NPs. A silica (SiO2) layer 5 nm thick coating evaporated Ag NPs with an average size of 60 nm is deposited on top of the zinc oxide (ZnO) layer used as the electron transport layer, leading to a significant improvement in the short‐circuit current density (Jsc) and the power conversion efficiency (PCE) of the inverted OSCs. The electron‐beam evaporation method is employed for controlled deposition of Ag NPs and SiO2 on the ZnO layer. The plasmonic devices resulted in an 18% and 14.1% enhancement of the Jsc and PCE, respectively, compared to reference devices. This increase of the photoelectric parameters in plasmonic devices is attributed not only to the plasmonic effects originating from the Ag NPs but also to the ultrathin silica layer which can contribute to facilitating charge extraction. © 2019 Society of Chemical Industry  相似文献   

17.
A catalytic method employing the cationic iridium‐(Sc,Rp)‐DuanPhos [(1R,1′R,2S,2′S)‐2,2′‐di‐tert‐butyl‐2,2′,3,3‐tetrahydro‐1H,1′H‐1,1′‐biisophosphindole] complex and BARF {tetrakis[3,5‐bis(trifluoromethyl)phenyl]borate} counterion effectively catalyzes the enantioselective hydrogenation of acyclic N‐arylimines with high turnover numbers (up to 10,000 TON) and excellent enantioselectivities (up to 98% ee), achieving the practical synthesis of chiral secondary amines.  相似文献   

18.
A catalytic complex made from [Ir(COD)Cl]2 [di‐μ‐chloro‐bis(1,5‐cyclooctadiene)diiridium(I)] precursor and (S,S)‐f‐Binaphane ((R,R)‐1,1′‐bis{(R)‐4,5‐dihydro‐3H‐dinaphtho[1,2‐c:2′,1′‐e]phosphepino}ferrocene) ligand effectively catalyzed the enantioselective hydrogenation of cyclic imines with high reactivity and good enantioselectivity.  相似文献   

19.
Since the first definition of domino reactions by Tietze in 1993, an explosive number of these fascinating reactions has been developed, allowing the easily building of complex chiral molecular architectures from simple materials to be achieved in a single step. Even more interesting, the possibility to join two or more reactions in one asymmetric domino process catalyzed by chiral metal catalysts has rapidly become one challenging goal for chemists, due to economical advantages, such as avoiding costly protecting groups and time‐consuming purification procedures after each step. The explosive development of enantioselective metal‐catalyzed domino including multicomponent reactions is a consequence of the considerable impact of the advent of asymmetric transition metal catalysis. This review aims to update the last developments of enantioselective one‐, two‐ and multicomponent domino reactions mediated by chiral metal catalysts, covering the literature since the beginning of 2006. Abbreviations: Ac: acetyl; AQN: anthraquinone; Ar: aryl; bdpp: 2,4‐bis(diphenylphosphino)pentane; BINAP: 2,2′‐bis(diphenylphosphino)‐1,1′‐binaphthyl; BINEPINE: phenylbinaphthophosphepine; BINIM: binapthyldiimine; BINOL: 1,1′‐bi‐2‐naphthol; BIPHEP: 2,2′‐bis(diphenylphosphino)‐1,1′‐biphenyl; Bn: benzyl; Boc: tert‐butoxycarbonyl; Box: bisoxazoline; BOXAX: 2,2′‐bis(oxazolyl)‐1,1′‐binaphthyl; BPTV: N‐benzene‐fused phthaloyl‐valine; Bu: butyl; Bz: benzoyl; Cat: catechol; Chiraphos: 2,3‐bis(diphenylphosphine)butane; cod: cyclooctadiene; Cy: cyclohexyl; DABCO: 1,4‐diazabicyclo[2.2.2]octane; dba: (E,E)‐dibenzylideneacetone; DBU: 1,8‐diazabicyclo[5.4.0]undec‐7‐ene; DCE: dichloroethane; de: diastereomeric excess; DHQ: hydroquinine; DHQD: dihydroquinidine; DIFLUORPHOS: 5,5′‐bis(diphenylphosphino)‐2,2,2′,2′‐tetrafluoro‐4,4′‐bi‐1,3‐benzodioxole; DIPEA: diisopropylethylamine; DMF: dimethylformamide; DMSO: dimethyl sulfoxide; DOSP: Np‐dodecylbenzenesulfonylprolinate; DPEN: 1,2‐diphenylethylenediamine; dtb: di‐tert‐butyl; dtbm: di‐tert‐butylmethoxy; E: electrophile; ee: enantiomeric excess; Et: ethyl; FBIP: ferrocene bis‐imidazoline bis‐palladacycle; Fc: ferrocenyl; FOXAP: ferrocenyloxazolinylphosphine; Hex: hexyl; HFIP: hexafluoroisopropyl alcohol; HMPA: hexamethylphosphoramide; iPr‐DuPhos: 1,2‐bis(2,5‐diisopropylphospholano)benzene; Josiphos: 1‐[2‐(diphenylphosphino)ferrocenyl]ethyldicyclohexylphosphine ethanol adduct; L: ligand; MCPBA: 3‐chloroperoxybenzoic acid; Me: methyl; Me‐DuPhos: 1,2‐bis(2,5‐dimethylphospholano)benzene; MEDAM: bis(dimethylanisyl)methyl; MOM: methoxymethyl; Naph: naphthyl; NMI: N‐methylimidazole; MWI: microwave irradiation; Norphos: 2,3‐bis(diphenylphosphino)‐bicyclo[2.2.1]hept‐5‐ene; Ns: nosyl (4‐nitrobenzene sulfonyl); Nu: nucleophile; Oct: octyl; Pent: pentyl; Ph: phenyl; PHAL: 1,4‐phthalazinediyl; Pin: pinacolato; PINAP: 4‐[2‐(diphenylphosphino)‐1‐naphthalenyl]‐N‐[1‐phenylethyl]‐1‐phthalazinamine; Pr: propyl; Py: pyridyl; PYBOX: 2,6‐bis(2‐oxazolyl)pyridine; QUINAP: 1‐(2‐diphenylphosphino‐1‐naphthyl)isoquinoline; QUOX: quinoline‐oxazoline; Segphos: 5,5′‐bis(diphenylphosphino)‐4,4′‐bi‐1,3‐benzodioxole; Solphos: 7,7′‐bis(diphenylphosphino)‐3,3′,4,4′‐tetrahydro‐4,4′‐dimethyl‐8,8′‐bis‐2H‐1,4‐benzoxazine; SPRIX: spirobis(isoxazoline); SYNPHOS: 6,6′‐bis(diphenylphosphino)‐2,2′,3,3′‐tetrahydro‐5,5′‐bi‐1,4‐benzodioxin; Taniaphos: [2‐diphenylphosphinoferrocenyl](N,N‐dimethylamino)(2‐diphenylphosphinophenyl)methane; TBS: tert‐butyldimethylsilyl; TC: thiophene carboxylate; TCPTTL: N‐tetrachlorophthaloyl‐tert‐leucinate; TEA: triethylamine; Tf: trifluoromethanesulfonyl; TFA: trifluoroacetic acid; THF: tetrahydrofuran; TMS: trimethylsilyl; Tol: tolyl; Ts: 4‐toluenesulfonyl (tosyl); C3‐Tunephos: 1,13‐bis(diphenylphosphino)‐7,8‐dihydro‐6H‐dibenzo[f,h][1,5]dioxonin; VAPOL: 2,2′‐diphenyl‐[3,3′‐biphenanthrene]‐4,4′‐diol  相似文献   

20.
Stille polymerization was employed to synthesize a low-band-gap (LBG) conjugated main-chain polymer PBTH consisting of bithiazole, dithieno[3,2-b:2′,3′-d]pyrroles (DTP), and pendent melamine derivatives. Novel supramolecular polymer networks PBTH/C and PBTH/F were developed by mixing proper molar amounts of polymer PBTH (containing melamine pendants) to be hydrogen-bonded (H-bonded) with complementary uracil-based conjugated cross-linkers C and F (i.e., containing two symmetrical uracil moieties connected with carbazole and fluorene units through triple bonds). The formation of multiple H-bonds between polymer PBTH and cross-linkers C or F was confirmed by FT-IR measurements. In contrast to polymer PBTH, the supramolecular design with multiple H-bonds can enhance the photovoltaic properties of polymer solar cell (PSC) devices containing H-bonded polymer networks PBTH/C and PBTH/F by tuning their light harvesting capabilities, HOMO energy levels, and crystallinities. Initially, the power conversion efficiency (PCE) values of PSC devices containing supramolecular polymer networks PBTH/C and PBTH/F as electron donors and [6,6]-phenyl-C71-butyric acid methyl ester (PC70BM) as an electron acceptor (polymer:PC70BM = 1:1 w/w) are found to be 0.97 and 0.68%, respectively, in contrast to 0.52% for polymer PBTH. The highest PCE value of 1.56% with a short-circuit current densities (Jsc) value of 7.16 mA/cm2, a open circuit voltages (Voc) value of 0.60 V, and a fill factor (FF) of 0.36 was further optimized in the PSC device containing a supramolecular polymer network PBTH/C as polymer:PC70BM = 1:2 w/w. These results indicate that supramolecular design is an effective route towards better photovoltaic properties of Voc, Jsc, and PCE values in polymer solar cells.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号