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 46 毫秒
1.
The construction and precise control of the face‐to‐face π‐stacked arrangements of anthracene fluorophores in the crystalline state led to a remarkable red shift in the fluorescence spectrum due to unprecedented excited oligomer formation. The arrangements were regulated by using organic salts including anthracene‐1,5‐disulfonic acid (1,5‐ADS) and a variety of aliphatic amines. Because of the smaller number of hydrogen atoms at the edge positions and the steric effect of the sulfonate groups, 1,5‐ADS should prefer face‐to‐face π‐stacked arrangements over the usual edge‐to‐face herringbone arrangement. Indeed, as the alkyl substituents were lengthened, the organic salts altered their anthracene arrangement to give two‐dimensional (2D) edge‐to‐face and end‐to‐face herringbone arrangements, one‐dimensional (1D) face‐to‐face zigzag and slipped stacking arrangements, a lateral 1D face‐to‐face arrangement like part of a brick wall, and a discrete monomer arrangement. The monomer arrangement behaved as a dilute solution even in the close‐packed solid state to emit deep blue light. The 1D face‐to‐face zigzag and slipped stacking of the anthracene fluorophores caused a red shift of 30–40 nm in the fluorescence emission with respect to the discrete arrangement, probably owing to ground‐state associations. On the other hand, the 2D end‐to‐face stacking induced a larger red shift of 60 nm, which is attributed to the excimer fluorescence. Surprisingly, the brick‐like lateral face‐to‐face arrangement afforded a remarkable red shift of 150 nm to give yellow fluorescence. This anomalous red shift is probably due to excited oligomer formation in such a lateral 1D arrangement according to the long fluorescence lifetime and little shift in the excitation spectrum. The regulation of the π‐stacked arrangement of anthracene fluorophores enabled the wide modulation of the fluorescence and a detailed investigation of the relationships between the photophysical properties and the arrangements.  相似文献   

2.
Radical ring‐opening polymerization of cyclic ketene acetals is a means to achieve novel types of aliphatic polyesters. 2‐methylene‐1,3‐dioxe‐5‐pene is a seven‐membered cyclic ketene acetal containing an unsaturation in the 5‐position in the ring structure. The double bond functionality enables further reactions subsequent to polymerization. The monomer 2‐methylene‐1,3‐dioxe‐5‐pene was synthesized and polymerized in bulk by free radical polymerization at different temperatures, to determine the structure of the products and propose a reaction mechanism. The reaction mechanism is dependent on the reaction temperature. At higher temperatures, ring‐opening takes place to a great extent followed by a new cyclization process to form the stable five‐membered cyclic ester 3‐vinyl‐1,4‐butyrolactone as the main reaction product. Thereby, propagation is suppressed and only small amounts of other oligomeric products are formed. At lower temperatures, the cyclic ester formation is reduced and oligomeric products containing both ring‐opened and ring‐retained repeating units are produced at higher yield. © 2009 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 47: 4587–4601, 2009  相似文献   

3.
The synthesis of linear and branched oligothiophenes of well‐defined structures is performed with regioselective deprotonation of 3‐substituted thiophenes and nickel‐catalyzed cross‐coupling of the thus formed metalated species with a bromothiophene. The reaction of 3‐hexylthiophene with EtMgCl and 2,2,6,6‐tetramethylpiperidine (TMP‐H, 10 mol %) induces the metalation selectively at the 5‐position by use of the catalytically generated hindered magnesium amide (TMPMgCl) and the subsequent reaction of a 2‐halo‐3‐hexylthiophene (bromide or chloride) in the presence of a nickel catalyst affords a head‐to‐tail (HT)‐type dimer. By repeating the same sequence, the linear oligothiophene up to a 4‐mer is synthesized in good yield. The reaction of 3‐hexylthiophene with 2,3‐dibromothiophene also takes place to afford a branched oligothiophene 3‐mer in quantitative yield. The obtained 3‐mer is also metalated at the sterically less‐hindered position in a regioselective manner furnishing a 7‐mer in >99 % yield after a further coupling reaction with 2,3‐dibromothiophene. These dendrimers react with several multifunctionalized organic electrophiles, leading to a variety of branched oligothiophenes.  相似文献   

4.
The unimolecular gas‐phase elimination kinetics of 2‐methoxy‐1‐chloroethane, 3‐methoxy‐1‐chloropropane, and 4‐methoxyl‐1‐chloroburane has been studied by using density functional theory (DFT) methods to propose the most reasonable mechanisms of decomposition of the aforementioned compounds. Calculation results of 2‐methoxy‐1‐chloroethane and 3‐methoxy‐1‐chloropropane suggest dehydrochlorination through a concerted nonsynchronous four‐centered cyclic transition state (TS) to give the corresponding olefin. In the case of 4‐methoxyl‐1‐chloroburane, in addition to the 1,2‐elimination mechanism, the anchimeric assistance by the methoxy group, through a polar five‐centered cyclic TS, provides additional pathways to give 4‐methoxy‐butene, tetrahydrofuran and chloromethane. The bond polarization of the C? Cl, in the direction of Cδ+ ··· Clδ?, is the limiting step of these elimination reactions. The significant increase in rate together with the formation of a cyclic product tetrahydrofuran in the gas‐phase elimination of 4‐methoxyl‐1‐chloroburane is attributed to neighboring group participation of the oxygen of the methoxy group in the TS. The theoretical calculations show a good agreement with the reported experimental results. © 2011 Wiley Periodicals, Inc. Int J Quantum Chem, 2011  相似文献   

5.
The spontaneous reactions of 1‐(2,2‐dimethyl‐1,3‐dioxane‐4,6‐dione‐5‐ylidene)‐4‐(dicyanomethylene)‐2,5‐cyclohexadiene (QM‐1) with a vinyl ether, butyl vinyl ether (BVE), and a cyclic ketene acetal, 2‐methylene‐1,3‐dioxepane (MDOP), were investigated. The reaction of QM‐1 with BVE produced a terpolymer composed of QM‐1, 7‐butoxy‐8,8‐dicyanoquinodimethane, and BVE units as a hexane‐insoluble product and a one‐to‐one adduct of methylene Meldrum's acid and BVE as a hexane‐soluble product. The spontaneous reaction of QM‐1 with BVE produced, in the presence of 2,2,6,6‐tetramethylpiperidine‐1‐oxy (TEMPO), a terpolymer carrying TEMPO units in the chain ends, and in the presence of methanol, a one‐to‐one‐to‐one adduct of QM‐1, BVE, and methanol was isolated. The spontaneous reaction with bulkier, electron‐donating MDOP produced a low‐molecular‐weight alternating cooligomer of QM‐1 with MDOP. The spontaneous polymerization was proposed to proceed via a zwitterionic intermediate taking two forms, gauche and trans, depending on the bulkiness of the comonomer. © 2004 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 42: 3800–3811, 2004  相似文献   

6.
刘河  刘春河  吴波  仲伯华  宫泽辉 《中国化学》2005,23(12):1589-1592
The reaction of 7α-acetyl-6,14-endoethano-6,7,8,14-tetrahydrothebaine with 2-(thien-2-yl)ethylmagnesium bromide was investigated. The tertiary alcohol derivative 7α-[R-l-hydroxyl-l-methyl-3-(thien-2-yl)propyl])-6,14- endoethano-6,7,8,14-tetrahydrothebaine (3) and a by-product 4 were isolated. The structure of 4 was elucidated by X-ray analysis. The Grignard reaction shows high degree of stereoselectivity according with Cram rule. The crystal structure of 4 indicates that dihydrofuran ring was opened to form a phenolic hydroxyl group and a three-membered ring structure. It maintains the main rigid structure of morphine and contains a C(6)-C(14) enthano bridge. The 1-hydroxyl-1-methyl-3-(thien-2-yl)propyl group at C(7) position adopted S-configuration.  相似文献   

7.
The reactions of 4‐(methylamino)‐3‐penten‐2‐one with diazoketones yielded the α‐acylenaminoketones 1–3 in good yields. Preparation of the α‐acylenaminoketone 4 was carried out by treatment of 4‐(t‐butyl‐amino)‐3‐penten‐2‐one with benzoyl chloride being followed by reaction of transamination with methyl‐amine. The reactions were carried out in five different solvents and were submitted to gas chromatogra‐phy/mass spectrometry analysis, with the goal of obtaining substituted pyrazoles and determining which of the carbonyls would preferentially be attacked by the nucleophile. The reactions of compounds 1–4 with hydrazine reagents led to the formation of the pyrazoles 5–7a‐q . Small amounts of 4‐methylamino‐2‐pentenones 10a‐q , amides 11a‐q and pyrazoles 12a‐q were also obtained in these reactions. The unexpected formation of pyrazoles 15d,h,q was detected when methanol and N,N‐dimethylformamide were used as solvents in the reactions of α‐acylenaminoketone 4 with hydrazine reagents.  相似文献   

8.
The first total synthesis of Bauerine C, a unique indoloquinazoline alkoloid, has been achieved from readily available 2,3‐dichloroaniline. The key step is the Japp–Klingmann condensation between 2,3‐dichloroaniline and ethyl‐2‐acetyl‐5‐phthalimido pentanoate to get 3‐[(2,3‐dichlorophenyl)‐hydrozono]‐pipiridin‐2‐one, which cyclizes to 7,8‐dichloro‐2,3,4,9‐tetrahydro‐β‐carbolin‐1‐one, which can be methylated by dimethyl sulphate to give 7,8‐dichloro‐9‐methyl 2,3,4,9‐tetrahydro‐β‐carbolin‐1‐one. This N‐methyl derivative is then subjected to dehydrogenation with 2,3‐dichloro‐5,6‐dicayano‐1,4‐benzoquinone (DDQ) to give the target compound Bauerine C.  相似文献   

9.
Aggregation‐induced emission (AIE) is commonly observed for propeller‐like luminogens with aromatic rotors and stators. Herein, we report that a coumarin derivative containing a seven‐membered aliphatic ring (CD‐7) but no rotors showed typical AIE characteristics, whereas its analogue with a five‐membered aliphatic ring (CD‐5) exhibited an opposite aggregation‐caused quenching (ACQ) effect. Experimental and theoretical results revealed that a large aliphatic ring in CD‐7 weakens structural rigidity and promotes out‐of‐plane twisting of the molecular backbone to drastically accelerate nonradiative excited‐state decay, thus resulting in poor emission in solution. The restriction of twisting motion in aggregates blocks the nonradiative decay channels and enables CD‐7 to fluoresce strongly. The results also show that AIE is a general phenomenon and not peculiar to propeller‐like molecules. The AIE and ACQ effects can be switched readily by the modulation of molecular rigidity.  相似文献   

10.
Corrosion rates are influenced by the formation of inhibitor aggregates on the copper surface. Laser scanning confocal microscope was used to investigate the adsorbed structures of benzotriazole, N‐((6‐methyl‐1H‐benzo[d][1,2,3]triazol‐1‐yl)methyl)‐N‐octyloctan‐1‐amine (MBA) and 2,5‐bis (ethyldisulfanyl)‐1,3,4‐thiadiazole at copper surface in relation to their performance as a corrosion inhibitor. The increase of water contact angle in the presence of inhibitor indicates its strong adsorption to the copper, and laser scanning confocal microscope visualization confirms the formation of MBA aggregates. The aggregates change from hemispherical to cylindrical shape with MBA concentration increasing in rolling oil, resulting in a decrease in corrosion rates as determined by mass loss measurements. Compared with 2,5‐bis (ethyldisulfanyl)‐1,3,4‐thiadiazole, oil‐added MBA has a quicker adsorption and formation speed of cylindrical aggregates and a lower corrosion rate. The adsorption of inhibitors on copper surface obeys Langmuir isotherm and physisorption and chemisorption mechanism. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

11.
杨新丽  乔丽明  戴维林 《催化学报》2015,(11):1875-1885
磷钨酸具有酸性,而且具有氧化还原性,是一种多功能的新型催化剂,具有很高的催化活性,稳定性好,既可作均相催化剂,也可做多相催化剂.磷钨酸作为多相催化剂主要负载于无机氧化物、介孔分子筛、活性炭和离子交换树脂等材料中,然而这些多相催化剂存在着结构不明确,磷钨酸分散不均、易流失、活性点易中毒等问题.为了克服以上问题,需要寻找更加合适的载体来制备新颖的负载型的磷钨酸多相催化剂,金属-有机骨架的独特性质,使该材料成为一种优良的催化剂载体.金属-有机骨架(MOFs)又称配位聚合物,是指由金属离子与有机配体通过配位键和其他一些弱作用力连接而成的具有超分子微孔网络结构的一种颇具应用前途的类沸石材料.这种材料具有丰富的孔结构和很大的比表面积,同时具有孔结构规整、孔径大小设计可调、表面化学基团修饰可调等优点,使得它在吸附分离、多相催化、环境保护等领域具有很好的应用前景. HKUST-1(Cu-BTC或MOF-199)是该领域内研究和应用较多的一种金属-有机骨架材料,它最早由香港大学Williams教授课题组报道,其为面心立方晶体.在结构中,每个Cu2簇与四个均苯三甲酸相连,每个均苯三甲酸桥连着三个Cu2簇,形成轮浆式次级结构单元.这些次级结构单元相互交错连接形成3D网络结构,具有孔径约为0.9 nm ×0.9 nm的正方形孔道,孔道中的客体分子可以除去,并可以为其它的客体分子所置换. HKUST-1本身即是一种优良的催化剂,同时也可作为一种性能稳定的催化剂载体.目前,关于HKUST-1在催化领域中的应用主要限制在微孔范围,其较小的孔道不利于物质扩散和传输,从而限制其实际应用.本论文利用超分子模板法,以十六烷基三甲基溴化铵(CTAB)为模板剂、铜为金属中心、均苯三甲酸为有机配体、磷钨酸(HPWs)为活性组分,采用一步水热法合成微-介孔多级孔道金属-有机骨架固载磷钨酸催化剂HPWs@Meso-HKUST-1,详细研究了该催化剂对环戊烯选择氧化制备戊二醛的催化性能,并采用X射线粉末衍射(XRD)、傅里叶红外光谱(FT-IR)、N2吸附、透射电镜(TEM)和室温CO原位吸附红外(CO-FT-IR)等表征手段对HPWs@Meso-HKUST-1催化剂进行了结构表征,从而解释该催化剂对目标反应具有优良催化性能的本质原因. N2吸附表征结果说明, HPWs@Meso-HKUST-1催化剂的吸附-脱附曲线在低相对压力范围内呈现I型吸附等温线,在高相对压力范围内呈现具有H2型滞后环的IV型吸附等温线;催化剂独特的吸附等温线表明以CTAB为模板剂,采用一步水热合成法可以得到具有微孔和介孔多级孔道的催化剂材料.催化剂的比表面积和孔容随着磷钨酸含量的增加而减少,结合文献报道,可以得出一步水热合成法使活性组分HPWs分布在载体的介孔孔道内. XRD和FT-IR测试结果表明,一步水热合成法可以成功的将HPWs引入HKUST-1中,且HPWs高度分散在载体中; HPWs@Meso-HKUST-1催化剂保持了载体HKUST-1的骨架结构.小角XRD和TEM结果说明,催化剂的多级孔结构为无序蠕虫状介孔组织.室温CO-FT-IR说明,在HPWs@Meso-HKUST-1催化剂中, HPWs提供了不同于载体HKUST-1的L酸酸性位.从以上结果可以得出,一步水热合成法使HPWs包裹在载体的介孔孔道内,防止了HPWs的流失,使HPWs@Meso-HKUST-1催化剂为环戊烯选择氧化制备戊二醛提供了大量的、高度分散的、具有L酸酸性位的活性中心,且催化剂的介孔孔道有利于反应物和产物的扩散,从而使该催化剂表现出优良的催化性能;在优化条件下,环戊烯的转化率达到92.5%,戊二醛的得率达到78.9%;热过滤实验表明该催化剂是真正的多相催化剂,且至少可以重复使用3次.  相似文献   

12.
Starting from commonly available sugar derivatives, a single step protocol to access a small family of isohexide‐dioxalates ( 2a–c ) has been established. The synthetic competence of 2a–c has been demonstrated by subjecting them to condensation polymerization. Quite surprisingly, the proton NMR of poly(isomannide‐co‐hexane)oxalate revealed a 1:2 ratio between isomannide‐dioxalate ( 2a ) and 1,6‐hexanediol ( 3a ) in the polymer backbone. This intriguing reactivity was found to be an outcome of a cross metathesis reaction between 2a and 3a . The cross metathesis products 3a ”[2‐(2‐methoxyacetoxy)ethyl 2‐(2‐hydroxyethoxy)‐2‐(λ3‐oxydanylidene)acetate] and 2a ‘(3R,6R)‐6‐hydroxyhexahydrofuro[3,2‐b]‐furan‐3‐yl methyl oxalate were isolated in a control experiment. Based on direct and indirect evidence, and control experiments, an alternative polymerization mechanism is proposed. Polymerization conditions were optimized to obtain polyoxalates P1(2a‐3a)‐P9(2c‐3c) with molecular weights in the range of 14,000–68,000 g/mol, and narrow polydispersities. The identity of the polyoxalates was unambiguously established using 1‐2D NMR spectroscopy, MALDI‐ToF‐MS, and GPC measurements. The practical implication of these polymers is demonstrated by preparing transparent, mechanically robust films. The environmental footprint of the selected polyoxalates was investigated by subjecting them to solution and solid‐state degradation. The polyoxalates were found to be amenable to degradation. © 2018 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2018 , 56, 1584–1592  相似文献   

13.
The syntheses of poly(1,3‐dioxan‐5‐yl methacrylate), poly(cis‐2‐phenyl‐1,3‐dioxan‐5‐yl methacrylate), poly(trans‐2‐phenyl‐1,3‐dioxan‐5‐yl methacrylate), poly(cis‐2‐cyclohexyl‐1,3‐dioxan‐5‐yl methacrylate), and poly(trans‐2‐cyclohexyl‐1,3‐dioxan‐5‐yl methacrylate) are reported. The mechanical relaxation spectrum of the simplest polymer, poly(1,3‐dioxan‐5‐yl methacrylate), exhibits a prominent β relaxation centered at ?98 °C, at 1 Hz, followed in increasing order of temperature by an ostensible glass–rubber relaxation process. In addition to the β relaxation, the loss curves of poly(trans‐2‐phenyl‐1,3‐dioxan‐5‐yl methacrylate) and poly(trans‐2‐cyclohexyl‐1,3‐dioxan‐5‐yl methacrylate) display in the glassy state a high activation energy relaxation, named the β* process, that seems to be a precursor of the glass–rubber relaxation of these polymers. The mechanical spectra of poly(trans‐2‐cyclohexyl‐1,3‐dioxan‐5‐yl methacrylate) and poly(cis‐2‐cyclohexyl‐1,3‐dioxan‐5‐yl methacrylate) exhibit a low activation energy process in the low‐temperature side of the spectra, which is absent in the other polymers. The molecular origin of the mechanical activity of these polymers in the glassy state is discussed in qualitative terms. © 2002 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 40: 1154–1162, 2002  相似文献   

14.
2‐Thioxo‐1,2,5,6,7,8‐hexahydroquinoline‐3‐carbonitrile ( 2 ) was easily S‐alkylated to produce alkyl mercapto derivatives 3a‐g . The latter compounds were cyclized to afford thienotetrahydroquinolines 4a‐g . Several pyrimidothienotetrahydroquinolines 5a‐d , and 6a‐d were obtained from the condensation of compounds 4c‐f with different reagents. o‐Aminocarbohydrazide derivative 11 was reacted with aromatic aldehydes, acetylacetone, nitrous acid and CS2 to afford compounds 12–15 . Compound 24 was coupled with aryldiazonium chloride to afford arylazo derivatives 25 . Also it condensed with aromatic aldehydes to give arylidene derivatives 26 . The latter compounds were reacted with malononitrile to give pyrano derivative 27 .  相似文献   

15.
The anthraquinone derivatives T‐x‐x ( x = 2, 4, and 8), possessing two cellobiosyl, cellotetraosyl, and cellooctaosyl chains, respectively, C‐glycosidically bonded at C(1) and C(8) were synthesised as potential mimics of cellulose I. The anthraquinone template enforces a parallel orientation of the cellodextrin chains at a distance corresponding to the one between the crystallographically independent chains of cellulose I, and the ethynyl and buta‐1,3‐diynyl linker units ensure an appropriate phase shift between them. The H‐bonding of the T‐x‐x mimics was analysed and compared to the one of the mono‐chained analogues T‐x and of the known cellulose II mimics N‐x‐x and N‐x where one or two cellodextrin chains are O‐glycosidically bonded to naphthalene‐1,8‐diethanol, or to naphthalene‐1‐ethanol. The OH signals of T‐x and T‐x‐x in solution in (D6)DMSO were assigned on the basis of DQFCOSY, HSQC, and TOCSY (only of T‐4, T‐4‐4 , and T‐8‐8 ) spectra and on a comparison with the spectra of N‐x and N‐x‐x. Hydrogen bonding was analysed on the basis of the chemical shift of OH groups and its temperature dependence, coupling constants, SIMPLE 1H‐NMR experiments, and ROESY spectra. T‐4‐4 and T‐8‐8 in (D6)DMSO appear to adopt a V‐shape arrangement of the cellosyl chains, avoiding inter‐chain H‐bond interactions. The well‐resolved solid‐state CP/MAS 13C‐NMR spectra of the mono‐chained T‐x ( x = 1, 2, 4, and 8) show that only T‐8 is a close mimic of cellulose II. While the solid‐state CP/MAS 13C‐NMR spectrum of the C1‐symmetric diglucoside T‐1‐1 is well‐resolved, the spectra of T‐2‐2 and T‐4‐4 show broad signals, and that of T‐8‐8 is rather well resolved. The spectrum of T‐8‐8 resembles that of cellulose Iβ. A comparison of the X‐ray powder‐diffraction spectra of T‐8‐8 and T‐8 with those of celluloses confirms that T‐8‐8 is a H‐bond mimic of cellulose I and T‐8 one of cellulose II. Surprisingly, there is little difference between the CP/MAS 13C‐NMR spectra of the acetyl protected mono‐chained C‐glycosylated anthraquinone derivatives A‐x and the double‐chained A‐x‐x ( x = 2, 4, and 8). The spectra of A‐4 and A‐4‐4 resemble strongly the one of cellulose triacetate I ( CTA I ). The (less well‐resolved) spectra of the cellooctaosides A‐8 and A‐8‐8 , however, resemble the one of CTA II . The similarity between the solid‐state CP/MAS 13C‐NMR spectra of A‐4 and A‐4‐4 to the one of CTA I , and of A‐8 and A‐8‐8 to the one of CTA II is opposite to the observations in the acetylated cellodextrin series. The mono‐chained A‐x cellulose triacetate mimics 21 ( A‐2 ), 32 ( A‐4 ), and 55 ( A‐8 ) were synthesised by Sonogashira coupling of the cellooligosyl‐ethynes 15, 28 , and 50 , followed by selective deacetylation. Complete deacetylation provided the corresponding T‐x mimics. The double‐chained A‐x‐x mimics 24 ( A‐2‐2 ), 35 ( A‐4‐4 ), and 58 ( A‐8‐8 ) were prepared from A‐x by triflation and Sonogashira coupling with the cellosyl‐buta‐1,3‐diynes 19, 31 , and 53 . Their deacetylation provided the corresponding T‐x‐x mimics 25, 36 , and 59 . The cellooligosyl‐ethynes and cellooligosyl‐buta‐1,3‐diynes required for the Sonogashira coupling were prepared by stepwise glycosylation of the partially O‐benzylated β‐cellobiosyl‐ethyne and β‐cellobiosyl‐buta‐1,3‐diyne 13 and 17 , respectively, with the cellobiosyl donor 2 and the cellohexaosyl donor 47 .  相似文献   

16.
The polymerization of 9‐phenyl‐10,10‐dioxo‐thioxanthen‐9‐yl and 9‐phenyl‐10‐oxo‐9,10‐dihydroanthracen‐9‐yl methacrylates obtained under radical initiation (α,α‐azobisisobutyronitrile) in benzene solution proceeds with high isotactic specificity to afford homopolymers with a triad mm content higher than 95%, having presumably a helical main‐chain structure and showing significant resistance to solvolytic degradation in methanol. 9‐Phenyl‐10,10‐dipropyl‐9,10‐dihydroanthracen‐9‐yl methacrylate similarly affords isotactic polymers with an mm of 98% but is much less durable in contact with methanol. The high isotacticity observed for the aforementioned polymethacrylates as well as for poly(1‐phenyl‐dibenzosuberyl methacrylate), previously reported in the literature, reveal a tendency of ortho‐bridged triarylcarbinols to enforce isotacticity on their methacrylate polymers obtained under radical initiation. © 2001 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 39: 1180–1186, 2001  相似文献   

17.
To provide a reliable tool for investigating diffusion processes of the specific components of the human odor 3‐hydroxy‐3‐methylhexanoic acid and 3‐methyl‐3‐sulfanylhexan‐1‐ol through the snowpack, we developed and optimized an analytical method based on direct immersion solid‐phase microextraction followed by gas chromatography with mass spectrometry. Direct immersion solid‐phase microextraction was performed using polyacrylate fibers placed in aqueous solutions containing 3‐hydroxy‐3‐methylhexanoic acid and 3‐methyl‐3‐sulfanylhexan‐1‐ol. After optimization, absorption times of 120 min provided a good balance to shorten the analysis time and to obtain suitable amounts of extractable analytes. The extraction efficiency was improved by increasing the ionic strength of the solution. Although the absolute extraction efficiency ranged between 10 and 12% for 3‐hydroxy‐3‐methylhexanoic acid and 2–3% for 3‐methyl‐3‐sulfanylhexan‐1‐ol, this method was suitable for analyzing 3‐hydroxy‐3‐methylhexanoic acid and 3‐methyl‐3‐sulfanylhexan‐1‐ol concentrations of at least 0.04 and 0.20 ng/mL, respectively. The precision of the direct immersion solid‐phase microextraction method ranged between 8 and 16%. The variability within a batch of six fibers was 10–18%. The accuracy of the method provided values of 88–95 and 86–101% for 3‐hydroxy‐3‐methylhexanoic acid and 3‐methyl‐3‐sulfanylhexan‐1‐ol, respectively. The limit of detection (and quantification) was 0.01 ng/mL (0.04 ng/mL) for 3‐hydroxy‐3‐methylhexanoic acid and 0.06 ng/mL (0.20 ng/mL) for 3‐methyl‐3‐sulfanylhexan‐1‐ol. The signal versus concentration was linear for both compounds (r2 = 0.973–0.979). The stability of these two compounds showed that 3‐hydroxy‐3‐methylhexanoic acid was more stable in water than 3‐methyl‐3‐sulfanylhexan‐1‐ol. We applied the method to environmental samples in correspondence with an olfactory target buried previously.  相似文献   

18.
Asymmetric hetero‐Diels‐Alder (AHDA) reactions provide a multitude of opportunities for the highly efficient, regio‐ and stereoselective construction of various heterocycles in enantiomerically pure form. The asymmetric aza‐Diels‐Alder (A‐aza‐DA) reaction using diversely hetero‐dienophiles and hetero‐dienes have been increasingly developed as a valuable method for the synthesis of functionalized nitrogen ring systems. The purpose of this review is to give a detailed discussion of the A‐aza‐DA reaction particularly, the stereoselective reactions of imines as dienophiles with Dainshefsky dienes to obtain optically pure aza‐Diels‐Alder products. The development of stereoselective variants of the reaction make use of imines as the dienophile and Dainshefsky dienes is at the forefront of these studies. This review updates the A‐aza‐DA reactions covering the literature from 1972 till date  相似文献   

19.
The phenanthridinium chromophores 5‐ethyl‐6‐phenylphenanthridinium ( 1 ), 5‐ethyl‐6‐methylphenanthridinium ( 2 ), 3,8‐diamino‐5‐ethyl‐6‐methylphenanthridinium ( 3 ), and 3,8‐diamino‐5‐ethyl‐6‐(4‐N,N‐diethylaminophenyl)phenanthridinium ( 4 ) were characterized by their optical and redox properties. All dyes were applied in titration experiments with a random‐sequence 17mer DNA duplex and their binding affinities were determined. The results were compared to well‐known ethidium bromide ( E ). In general, this set of data allows the influence of substituents in positions 3, 6, and 8 on the optical properties of E to be elucidated. Especially, compound 4 was used to compare the weak electron‐donating character of the phenyl substituent at position 6 of E with the more electron‐donating 4‐N,N‐diethylaminophenyl group. Analysis of all of the measurements revealed two pairs of chromophores. The first pair, consisting of 1 and 2 , lacks the amino groups in positions 3 and 8, and, as a result, these dyes exhibit clearly altered optical and electrochemical properties compared with E . In the presence of DNA, a significant fluorescence quenching was observed. Their binding affinity to DNA is reduced by nearly one order of magnitude. The electronic effect of the phenyl group in position 6 on this type of dye is rather small. The properties of the second set, 3 and 4 , are similar to E due to the presence of the two strongly electron‐donating amino groups in positions 3 and 8. However, in contrast to 1 and 2 , the electron‐donating character of the substituent in position 6 of 3 and 4 is critical. The binding, as well as the fluorescence enhancement, is clearly related to the electron‐donating effect of this substituent. Accordingly, compound 4 shows the strongest binding affinity and the strongest fluorescence enhancement. Quantum chemical calculations reveal a general mechanism related to the twisted intramolecular charge transfer (TICT) model. Accordingly, an increase of the twist angle between the phenyl ring in position 6 and the phenanthridinium core opens a nonradiative channel in the excited state that depends on the electron‐donating character of the phenyl group. Access to this channel is hindered upon binding to DNA.  相似文献   

20.
To learn how a preorganized peptide‐based molecular template, together with diverse weak non‐covalent interactions, leads to an effective self‐association, we investigated the conformational characteristics of a simple γ,α‐hybrid model peptide, Boc‐γ‐Abz‐Gly‐OMe. The single‐crystal X‐ray diffraction analysis revealed the existence of a fully extended β‐strand‐like structure stabilized by two non‐conventional C?H???O=C intramolecular H‐bonds. The 2D 1H NMR ROESY experiment led us to propose that the flat topology of the urethane‐γ‐Abz‐amide moiety is predominantly preserved in a non‐polar environment. The self‐association of the energetically more favorable antiparallel β‐strand‐mimic in solid‐state engenders an unusual ‘flight of stairs’ fabricated through face‐to‐face and edge‐to‐edge Ar???Ar interactions. In conjunction with FT‐IR spectroscopic analysis in chloroform, we highlight that conformationally semi‐rigid γ‐Abz foldamer in appositely designed peptides may encourage unusual β‐strand or β‐sheet‐like self‐association and supramolecular organization stabilized via weak attractive forces.  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