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 15 毫秒
1.
Summary: Some peculiarities in the cyclopolymerization of 1,6-heptadiynes containing an amine-nitrogen at the 4-position are presented and discussed within the context of cyclopolymerization of N-free 1,6-heptadiynes.  相似文献   

2.
A fluorene‐containing spiro‐type conjugated polymer, poly(dipropargylfluorene), was synthesized via the cyclopolymerization of dipropargylfluorene by Mo‐ and W‐based transition‐metal catalysts. The polymerization of dipropargylfluorene proceeded well by MoCl5 catalyst itself to give a quantitative yield of polymer. The Mo‐based catalysts are more effective than those of W‐based catalysts. The structure of poly(dipropargylfluorene) was characterized by various instrumental methods (NMR, IR, and UV–visible spectroscopies) to have the conjugated polymer backbone carrying fluorene moieties. Analysis of the 13C NMR spectrum revealed that the polymer structure consists of only six‐membered rings. The resulting poly(dipropargylfluorene)s were brown or black powder and soluble in aromatics and halogenated hydrocarbons such as benzene, chlorobenzene, tetrahydrofuran, chloroform, and methylene chloride. Thermal and morphological properties of the polymer are also discussed. © 2001 John Wiley & Sons, Inc. J Polym Sci Part A: Polym Chem 39: 4101–4109, 2001  相似文献   

3.
A systematic study of the asymmetric cyclocopolymerization of bis(4‐vinylbenzoate)s, derived from chiral diols, with styrene has been made from the viewpoint of synthesizing the main‐chain chiral polymer. On the basis of using over 30 chiral diols as templates, we summarize the relationship between the structure of the chiral template and the chiroptical properties of the template‐free polymer. For simple chiral diols, the chirality induction efficiency increased in the order 1,2‐diol < 1,4‐diol < 1,3‐diol. Chiral diols with two chiral centers exhibited higher chirality induction efficiency than those having one chiral center only. The chirality induction efficiency for cyclic 1,2‐diols increased with the ring size in the order 5‐ < 6‐ < 7‐ < 8‐membered rings, and that for acyclic 1,2‐diols increased with the bulkiness of the substituent at the chiral center. In addition, a chirality induction mechanism has been proposed on the basis of model radical cyclization experiments and computational studies. Chirality induction could be caused by the inhibition of the formation of one racemo unit among the four stereoisomers due to the strong dependence of the stereoselectivity in intermolecular additions on the absolute configuration of the cyclized radical. The mechanism was examined using the Lewis‐acid and monomer‐concentration effects.  相似文献   

4.
Cyclopolymerization of 1,7-octadiynes using a ruthenium-based Grubbs catalyst, to produce conjugated polymers containing six-membered rings as repeat units is generally much slower than the corresponding polymerization of 1,6-heptadiynes, and thus it is considered less useful. Here, we demonstrate the regioselective cyclopolymerization of 4,5-disubstituted 1,7-octadiynes with considerably enhanced reactivity. Using a third generation Grubbs catalyst with a rapid initiation step, various conjugated polymers with low polydispersity indices (PDIs) could be synthesized under optimized conditions. Among the various monomers tested, those with bulky substituents underwent controlled polymerization within 1 h at room temperature, which was a significant improvement over previous reports. This led us to a more efficient preparation of fully conjugated block copolymers. Finally, owing to the fast cyclopolymerization, a synthetically challenging dendronized polymer was successfully prepared from a macromonomer containing two second generation dendrons at the 4 and 5 positions of 1,7-octadiyne, and its rod-like conformation was visualized using atomic force microscopy. © 2014 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2015 , 53, 274–279  相似文献   

5.
To develop the polymerization exploiting the interconversion of fluorinated carbon radical to hydrocarbon radical, the radical cyclopolymerization of perfluoroisopropenyl vinylacetate [CF2?C(CF3)OCOCH2CH?CH2] (FIA) was investigated to afford a polymer possessing mainly five‐membered ring structure with bimodal molecular weight distribution having 1 × 105 as the higher molecular weight. This may be the first example wherein the cyclopolymerization between usual allyl group and fluorinated vinyl group is performed. The degree of cyclization was between 70 and 80% determined by 19F NMR of as‐polymerized products. The polymer preparation from perfluoroisopropenyl group, which shows scarce homopolymerization reactivity was accomplished. The mechanism that the addition of hydrocarbon radical to perfluoroisopropenyl group to produce fluorinated carbon radical followed by the intramolecular addition reaction onto allyl group to form five‐membered ring is proposed. The hydrolysis of the FIA polymer afforded a polymer possessing hydrophobic fluoroalkyl group with hydrophilic hydroxyl and carboxylic acid groups. © 2006 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 44: 3220–3232, 2006  相似文献   

6.
In the past few decades, conjugated polymers have aroused extensive interest in organic electronic applications. The electrical performance of conjugated polymers has a close relationship with their backbone conformation. The conformation of the polymer backbone strongly affects the πelectron delocalization along polymer chains, the energy band gap, interchain interactions, and further affects charge transport properties. To realize a rigid coplanar backbone that usually possesses efficient intrachain charge transport properties and enhanced π–π stackings, such conformation control becomes a useful strategy to achieve high-performance (semi)conducting polymers. This minireview summarizes the most important polymer structures through conformation control at the molecular level, and then divides these rigid coplanar conjugated polymers into three categories: 1) noncovalent interactions locked conjugated polymers; 2) double-bond linked conjugated polymers; 3) ladder conjugated polymers. The effect of the conformation control on physical nature, optoelectronic properties, and their device performance is also discussed, as well as the challenges of chemical synthesis and structural characterization.  相似文献   

7.
An alternating copolymer of methacrylic acid and N‐isopropyl acrylamide (NIPAM) was synthesized by selective cyclopolymerization of a special divinyl monomer and transformation of repeating cyclo‐units in the resultant cyclopolymer. Crucial to the breakthrough is the monomer design in view of two types of cleavable bonds (3° ester and activated ester) in the pendant group of the monomer and the lower reactivity ratio of the two double bonds (methacrylate and electron‐poor acrylate) for the polymerizable groups. The thus‐obtained cyclopolymer was transformed into the alternating copolymer by transformation of the activated ester into amide by isopropyl amine and cleavage of the 3° ester by trifluoroacetic acid. The resultant copolymer showed thermal and pH response in aqueous solution that was different from the 1:1 random copolymer as well as the homopolymers.  相似文献   

8.
Novel ladder‐type conjugated polymers, fused poly (benzopentalene) derivatives, were synthesized from the readily accessible 1,4‐dibromo‐2,5‐diethynylbenzene derivatives by the Pd‐catalyzed self‐polycondensation in one‐step with high yields. The low solubility of the ladder structure was suggested when the triisopropylsilyl substituents were selected. However, when longer alkyl chains were introduced into the peripheral moieties, such as the dialkylanilino (DAA) and alkyloxyphenyl groups, a high solubility was achieved and the number‐average molecular weight (Mn) reached 18,000. The UV‐Vis absorption spectral shapes of the polymers were similar to the reported dibenzopentalene derivatives, except for the bathochromically shifted end absorptions. This result suggests an extension of the π‐conjugated systems due to the polymerization. Moreover, the almost defect‐free structure of the ladder‐type polymers was confirmed by the quantitative tetracyanoethylene (TCNE) addition to the DAA‐activated alkynes. The titration experiments of TCNE to the polymers revealed the number of terminal alkynes, which enabled us to calculate the molecular weight of the polymers. The calculated molecular weight was consistent with that determined by GPC. After the TCNE addition, the polymer band gaps reasonably decreased as suggested by the UV‐Vis‐NIR absorption and electrochemical measurements. © 2010 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2010  相似文献   

9.
A helical step‐ladder polyarylene incorporating chiral (R)‐2,2′‐dioctoxy‐1,1′‐binaphthyl units was synthesized for the first time. The first step involved the preparation of a precursor poly(arylene ketone) via a palladium‐mediated Suzuki‐type cross‐coupling reaction with the aid of microwave heating. Two polymer‐analog reaction steps, the reduction of the keto groups to tertiary alcohol functionalities and subsequent intramolecular Friedel–Crafts cyclization, gave a step‐ladder polymer ( 6 ) in good yields with reasonable mean average molecular weights greater than 13,000. The regioselective cyclization pattern in the α position of the naphthalene core was confirmed by a comparison of the NMR data of the polymer with those of the corresponding model ladder oligomers, 12 and 13 , and also a single‐crystal structure of 13 . The optical spectra of the oligomers and polymers indicated that there was little electronic interaction across the binaphthyl units. The circular dichroism spectrum of 6 exhibited a strong bisignate Cotton effect in the π–π* absorption region of the planar chromophores, which reflected the strong exciton coupling within the helical polymer chain. © 2006 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 44: 5533–5545, 2006  相似文献   

10.
The synthesis of two 1,3‐bis(4‐ethynylbenzyloxy)calix[4]arenes, 5,11,17,23‐tetrakis(1,1‐dimethylethyl)‐25,27‐bis(4‐ethynylbenzyloxy)‐26,28‐dihydroxycalix[4]arene ( 1 ) and 25,27‐bis(4‐ethynylbenzyloxy)‐26,28‐dihydroxycalix[4]arene ( 2 ), was accomplished through Sonogashira coupling of appropriate calixarene derivatives. Methods for the polymerization of these bifunctional building blocks with Rh(I) as a catalyst, leading ultimately to conjugated polymers having calix[4]arene units incorporated into the main chain, were explored. Calixarenes 1 and 2 were efficiently polymerized with rhodium‐based initiators and afforded the conjugated polymers poly{5,11,17,23‐tetrakis(1,1‐dimethylethyl)‐25,27‐bis(4‐ethynylbenzyloxy)‐26,28‐dihydroxycalix[4]arene} ( poly 1 ) and poly{25,27‐bis(4‐ethynylbenzyloxy)‐26,28‐dihydroxycalix[4]arene}. Depending on the conditions, high conversions and good yields were obtained. The effects of adding cocatalysts (NHEt2 and/or PPh3) were studied in connection with the number‐average molecular weight and the molecular weight distribution of the resultant polymer ( poly 1 ) and tentatively correlated with the formation of low‐molecular‐weight materials. A catalytic system containing triphenylphosphine as the sole additive ([Rh(nbd)Cl]2; [Rh]/[PPh3] = 0.5) proved to be the best for the polymerization of ptert‐butylcalixarene compound 1 . Linear polymers having high number‐average molecular weights (up to 1.1 × 105 g mol?1) with low polydispersities were produced under these conditions. For debutylated homologue 2 , its polymerization was best carried out in the absence of any added cocatalyst. A cyclopolymerization route, comprising the intramolecular ring closing of the calix[4]arene pendant ethynyl groups followed by an intermolecular propagation step, is advanced to explain the results. © 2006 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 44: 7054–7070, 2006  相似文献   

11.
Enantio‐ and stereoselective cyclopolymerization of hexa‐1,5‐diene was achieved by enantiomerically pure dichloro zirconium(IV) pre‐catalysts 2 possessing chiral [OSSO]‐type bis(phenolate) ligands (−)‐ 1 and (+)‐ 1 in combination with dried methylaluminoxane (dMAO) as an activator. The corresponding activities were recorded with quite high values up to 1,960 g mmol( 2 )–1 h–1, which are extremely larger than those of the related complexes. The microstructure analysis for the PMCPs furnished by pre‐catalysts (Λ,S,S)‐ 2 and (Δ,R,R)‐ 2 showed good isotacticity factors (α = 75−78%) and relatively high proportions of trans‐cyclopentane rings (σ = 14−21%). These enantiomeric PMCPs exhibited large specific optical rotations ([α]D = +28 to +32° from (Λ,S,S)‐ 2 , −26 to −34° from (Δ,R,R)‐ 2 ).

  相似文献   


12.
Two optically active solvents were synthesised, (S)‐(–)‐2‐methyl‐1‐propoxybutane and (S)‐(–)‐(2‐methylbutoxymethyl)benzene. The main chain conformations of poly(methylphenylsilane) and poly(hexylmethylsilane) in these solvents were investigated using optical UV‐visible and circular dichroism spectroscopy. It was observed that dissolving these inherently achiral polysilanes in optically active solvents induces the polymer chains to adopt preferred helical screw senses. This is the first example of induction of optical activity in conjugated polymers through chiral solvation.  相似文献   

13.
N,N-Diallylmethacrylamide (DAMA) was synthesized and its radical cyclopolymerizability was investigated. Polymerizations of DAMA were found to proceed according to the standard kinetic equation for radical polymerization except for the diffusion controlled termination reactions owing to higher viscosity of DAMA, in spite of the fact that it contains two allyl groups. DAMA yields polymers with slightly higher degree of cyclization than 50% in dilute solution polymerization. With increasing monomer concentration, the degree of cyclization decreases gradually and it becomes slightly lower value than 50% in concentrated solution polymerizations. It was found that about 86% of pendant unsaturations is allyl group and the rest is methacryloyl group. Structural studies on repeating units showed that possibility for the presence of bicyclic structures (those formed with consumption of three double bonds involved in DAMA) and structural units with two pendant unsaturations (those formed with one double bond in DAMA participated in polymerization) is extremely small. Main repeating cyclic unit was assigned to five-membered monocyclic lactam with an allyl group on its nitrogen, the content of which is estimated to be approximately 86% based on the allyl content. The main fraction of the rest of repeating cyclic structures was attributed to monocyclic pirolidine ring with a methacryloyl group on its nitrogen. These structural investigations and ESR studies of DAMA revealed that the methacryloyl group of DAMA is less reactive than its allyl group as in the case of N-methyl-N-allylmethacrylamide which has already been reported. © 1995 John Wiley & Sons, Inc.  相似文献   

14.
The effect of bulky N-substitutents of N-t-butyl- and N-phenyl-N-allylmethacrylamides (BAMA and PAMA, respectively) on their cyclopolymerizability was investigated. BAMA yielded an almost completely cyclized polymer while the degree of cyclization of poly (PAMA) was about 95%. The latter value indicates that the effect of phenyl group is comparable with that of methyl group, since N-methyl-N-allylmethacrylamide was reported to give a polymer with a degree of cyclization of 93%. Structural investigation on telomerization products of BAMA and PAMA permitted the assignment of the repeating cyclic units of these polymers to that of a five-membered ring. This structural characteristic was also supported by the observation of five-membered cyclized radicals on ESR measurements of these monomers. Rotation around amide C? N bonds of these monomers and related compounds studied by 13C-NMR was found to be strongly dependent on N-substitutents. The mechanism of the cyclization was discussed in terms of the structure of the ring formed and rotation around amide C? N bonds of these monomers. The reactivity of the methacryloyl and allyl groups involved in these monomers were compared based on the information obtained by structural investigation of polymers and telomerization products and by ESR studies. © 1993 John Wiley & Sons, Inc.  相似文献   

15.
Diacetylene monomers with respective lengths of the oxyethylene chains were cyclopolymerized with a rhodium catalyst to produce novel poly(phenylacetylene)s bearing a different cavity size of the chiral crown ether in the repeating units ( 2a – c ). In the circular dichroism spectra of the resulting polymers, characteristic Cotton effects were observed in the range from 350 to 500 nm corresponding to the absorption of the conjugated polymer backbone, indicating that the polymers possessed a helical structure with an excess single screw sense induced by the covalently bonded binaphthyl units. The host–guest complexation of 2a – c with achiral guests produced a chiroptical change based on the fluctuation in the main chain conformation. The behavior of the complexation‐induced chiroptical change was essentially dictated by the cavity size of the binaphthocrown ether units. Additionally, a chirality‐responsive helicity change was observed in the case of the complexation of 2a – c with chiral guests, which also depended on the crown ether size. © 2010 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 48: 1197–1206, 2010  相似文献   

16.
Chiral all-π-conjugated conducting polymers are of significant importance because of their promising aspects ranging from organic electronics, organic spin filters, efficient water splitting materials, magneto-optic materials to circularly polarized light emitters. Here, we report the synthesis and characterization of the first series of diketopyrrolopyrrole based chiral low band-gap alternating copolymers. Alongside characterizations by UV and CD spectroscopy, AFM and FESEM were further performed to understand the morphology of the optically active polymers. The stereochemistry of the side-chains on TDPP unit played a pivotal role in determining chiroptical properties and supramolecular self-assembly of the copolymers in solution which lead to layer-by-layer helical morphology in solid-state. Interestingly, D-A intramolecular charge transfer transition showed Cotton effect with bimodal nature, which is remarkably stronger than the π-π* transition.  相似文献   

17.
In this comment on the Essay entitled “Bad Language” by Dunitz I add new elements which enlighten the discussion. I hope that they contribute to suppress ambiguities and preconceived ideas about chirality.  相似文献   

18.
To discuss the difference between stereogenicity and chirality, we propose the concept of RS-stereoisomeric groups. Beginning with this concept, we have further proposed the concepts of holantimers, stereoisograms, and RS-stereogenicity. Thereby, we have clarified that the concept of RS-stereogenicity, but not conventional stereogenicity, is closely related to chirality. Thus, five RS-stereogenicity types are defined and examined to discuss the difference between stereogenicity and chirality. Combinatorial enumerations have also been studied by considering the RS-stereogenicity.  相似文献   

19.
Bad Language     
Bad language in stereochemistry—and elsewhere—can lead to sloppy thinking. In this Essay I review the history of stereochemical concepts and vocabulary in the hope that it may contribute a little to better thinking and communication.  相似文献   

20.
In this comment on the Essay entitled “Bad Language” by Dunitz I add new elements which enlighten the discussion. I hope that they contribute to suppress ambiguities and preconceived ideas about chirality.  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