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 31 毫秒
1.
Telechelic polyisobutylenes with silylcyclopentadiene termini (CpSi-PIB-SiCp) were prepared, starting with α,ω-di(t-chloro)polyisobutylene and followed by dehydrochlorination, hydrosilylation, and cyclopentadienylation. The prepolymer was characterized and found to have a terminal functionality close to 2.0. Thermal chain extension by Diels–Alder addition of the end groups did not occur, most likely because of fluxional isomerization of the silylcyclopentadiene end groups, confirmed by model studies. Chain coupling of CpSi-PIB-SiCp prepolymers with stoichiometric amounts of bismaleimide gave rise to degrees of extension of more than 11.  相似文献   

2.
Highly enantioselective Diels–Alder (DA) and inverse‐electron‐demand hetero‐Diels–Alder (HDA) reactions of β,γ‐unsaturated α‐ketoesters with cyclopentadiene catalyzed by chiral N,N′‐dioxide–Cu(OTf)2 (Tf=triflate) complexes have been developed. Quantitative conversion of β,γ‐unsaturated α‐ketoesters and excellent diastereoselectivities (up to 99:1) and enantioselectivities (up to >99 % ee) were observed for a broad range of substrates. Both aromatic and aliphatic β,γ‐unsaturated α‐ketoesters were found to be suitable substrates for the reactions. Moreover, the chemoselectivity of the DA and HDA adducts were improved by regulating the reaction temperature. Good to high chemoselectivity (up to 94 %) of the DA adducts were obtained at room temperature, and moderate chemoselectivity (up to 65 %) of the HDA adducts were achieved at low temperature. The reaction also featured mild reaction conditions, a simple procedure, and remarkably low catalyst loading (0.1–1.5 mol %). A strong positive nonlinear effect was observed.  相似文献   

3.
This article describes the synthesis and characterization of polyisobutylene (PIB) carrying one primary hydroxyl head group and a tertiary chloride end group, [Ph? C(CH3)(CH2OH)–PIB–CH2? C(CH3)2Cl] prepared with direct functionalization via initiation. The polymerization of isobutylene was initiated with the α‐methylstyrene epoxide/titanium tetrachloride system. Living conditions were obtained from ?75 to ?50 °C (198–223 K). Low molecular weight samples (number‐average molecular weight ~ 4000 g/mol) were prepared under suitable conditions and characterized by Fourier transform infrared and 1H NMR spectroscopy. The presence of primary hydroxyl head groups in PIB was verified by both methods. Quantitative Fourier transform infrared with 2‐phenyl‐1‐propanol calibration and 1H NMR performed on both the hydroxyl‐functionalized PIB and its reaction product with trimethylchlorosilane showed that each polymer chain carried one primary hydroxyl head group. The synthetic methodology presented here is an effective and simple route for the direct functionalization of PIB. © 2002 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 40: 1005–1015, 2002  相似文献   

4.
An efficient and short entry to polyfunctionalized linear triquinanes from 2‐methoxyphenols is described by utilizing the following chemistry. The Diels–Alder reactions of masked o‐benzoquinones, derived from 2‐methoxyphenols, with cyclopentadiene afford tricyclo[5.2.2.02,6]undeca‐4,10‐dien‐8‐ones. Photochemical oxa‐di‐π‐methane (ODPM) rearrangements and 1,3‐acyl shifts of the Diels–Alder adducts are investigated. The ODPM‐rearranged products are further converted to linear triquinanes by using an O‐stannyl ketyl fragmentation. Application of this efficient strategy to the total synthesis of (±)‐Δ9(12)‐capnellene was accomplished from 2‐methoxy‐4‐methylphenol in nine steps with 20 % overall yield.  相似文献   

5.
An organocatalyst formed from a binaphthyl‐substituted diamine and trifluoromethanesulfonic acid exhibited unprecedented levels of exo selectivity in the Diels–Alder reaction of α,β‐unsaturated aldehydes with cyclopentadiene. A novel axially chiral diamine was also designed as an organocatalyst for an asymmetric variant of this reaction, in which the desired cycloadducts were formed with high diastereo‐ and enantioselectivities. The highest diastereoselectivity observed was greater than 20:1 in favor of the exo cycloadduct in the asymmetric Diels–Alder reaction of crotonaldehyde with cyclopentadiene.  相似文献   

6.
This article presents the synthesis and initial characterization of polymers having a high molecular weight dendritic (arborescent) polyisobutylene (arbPIB) core and short end sequences composed of copolymers of isobutylene with isoprene, p‐methyl styrene, or cyclopentadiene. The first stage of the polymerizations was controlled/living (i.e., no chain transfer or irreversible termination were detected) and the process was reproducible. The hydrodynamic radii of arbPIBs were measured by viscometry and quasi‐elastic light scattering, and good agreement between the two methods is demonstrated for the first time in this paper. Sequential addition of the second monomer(s) led to the formation of short copolymer end sequences. We have discovered that regardless of the Tg of the copolymer end blocks, the new block copolymers exhibited TPE properties, which were improved by carbon black and silica fillers to a great extent. These novel PIB‐based TPEs are promising new biomaterials. © 2009 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 47: 1148–1158, 2009  相似文献   

7.
We have studied the solvent, temperature, and pressure influences on the reaction rates of cyclic and acyclic N=N bonds in the Diels–Alder and ene reactions. The transfer from N‐phenylmaleimide ( 9 ) to a structural analogue, 4‐phenyl‐1,2,4‐triazoline‐3,5‐dione ( 2 ), is accompanied by the rate increase in five to six orders of magnitude in the Diels–Alder reactions with cyclopentadiene ( 4 ) and 9,10‐dimethylanthracene ( 5 ), whereas the transfer from dimethyl fumarate ( 10 ) to diethyl azodicarboxylate ( 1 ) increases only in one to two orders of magnitude. The ratio of the reaction rate constants ( 2 + 4 )/( 1 + 4 ) is very large (5.2 × 107) and almost the same (5.3 × 107) as in the ene reactions with tetramethylethylene ( 7 ), ( 2 + 7 )/( 1 + 7 ). It has been observed that the N=N bond in reagent 2 has strong electrophilic, and its N–N moiety in the transition state has nucleophilic properties, which results from the analysis of the solvation enthalpy transfer of reagents, activated complex, and adduct in the Diels–Alder reaction of 2 with anthracene 22 .  相似文献   

8.
The synthesis and characterization of α,ω‐di(2‐methyl‐2,3‐epoxypropyl)polyisobutylene are reported. The epoxidation of α,ω‐di(isobutenyl)polyisobutylene was achieved at room temperature with dimethyldioxirane, which proved to be a very effective reagent for epoxidation without the formation of byproducts. A very good agreement was found for the conversion determined by 1H NMR and matrix‐assisted laser desorption/ionization mass spectrometry (MALDI HMS). The epoxy end groups were converted quantitatively into aldehyde termini with zinc bromide as a catalyst. The aldehyde groups were then reduced with LiAlH4 into primary hydroxyl functions to obtain α,ω‐di(2‐methyl‐3‐hydroxylpropyl)polyisobutylene with high efficiency. © 2002 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 40: 3974–3986, 2002  相似文献   

9.
A microporous polymer is prepared by a catalyst‐free Diels–Alder reaction. A cyclopentadiene with both a diene and a dienophile functionality and a dienophilic maleimide are used for the Diels–Alder reaction. 1,3,5‐Tris(bromomethyl)‐2,4,6‐trimethylbenzene is reacted with sodium cyclopentadienide to produce the multicyclopentadiene‐functionalized monomer. A crosslinked polymer ( CDAP ) is obtained by the reaction of the cyclopentadiene monomer with N,N′‐1,4‐phenylenedimaleimide. The thermal dissociation of the cyclopentadiene dimeric unit and the subsequent Diels–Alder reaction with the maleimide group are investigated by the model reaction. We are able to restructure the crosslinked polymer network by taking advantage of the thermal reversibility of the Diels–Alder linkage. After the post thermal treatment, the BET surface area of the polymer ( CDAP‐T ) is greatly increased from 317 to 1038 m2 g?1. CDAP‐T is functionalized with pyrene by bromination with N‐bromosuccinimide and the subsequent substitution reaction with aminopyrene. The adsorption property of the pyrene‐functionalized polymer for an aromatic dye is investigated using malachite green. © 2013 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2013, 51, 3646–3653  相似文献   

10.
Diels–Alder cycloaddition between cyclopentadiene and p-benzoquinone has been studied using amorphous silica and different ITQ-2 and MCM-41 pure silica and metal containing materials as catalysts. The reaction can afford different products depending on the molecular reacting site, and the possibility of consecutive additions. Structured solid catalysts increase the selectivity to the endo–endo isomer. Silanol groups have not enough Brönsted acidity to interact with the carbonyl groups present in the dienophile, to reduce LUMO's energy and provide a better overlap between HOMO and LUMO, according to the frontier molecular orbital theory.

The introduction of transition metal atoms in the framework increases the reaction rate for the Diels–Alder reaction while preserving the selectivity to the endo–endo isomer. The presence of more acidic OH groups enhance the retro-Diels–Alder reaction increasing the selectivity to the endo–exo isomer.  相似文献   


11.
《Tetrahedron: Asymmetry》2006,17(20):2943-2951
Enantiomerically pure allene-1,3-dicarboxylates were easily synthesized by using epimerization–crystallization of dissymmetric allene compounds, which were prepared from acetone-1,3-dicarboxylates and naturally abundant chiral alcohols, that is, (−)- and (+)-menthols, borneol, and isoborneol. After scrutinizing the crystallization of several allene-1,3-dicarboxylates in the presence of triethylamine, it was found that allene-1,3-dicarboxylate carrying bornyl groups was the most easily prepared as a single isomer because of it having suitable solubility to be crystallized in hexane at 0 °C to room temperature. Diels–Alder reaction of the enantiomerically pure allene-1,3-dicarboxylates and cyclic dienes, such as N-Boc-pyrrole and cyclopentadiene, afforded endo-adducts having the same configurations at two newly generated stereogenic centers.  相似文献   

12.
The synthesis of a poly(azine–ether) via Williamson etherification using the cesium salt of 4–hydroxyacetophenone azine and 1,10–dibromodecane was carried out in N-methyl–2–pyrrolidone. The heterogeneous reaction proceeded readily at temperatures from ambient to 150°C. Polymers of varying molecular weight with essentially alkyl bromide end groups were produced either by changing the polymerization temperature or by using an excess of the organic substrate. The thermal stability of the polymers was molecular weight dependent and those with the highest DPn exhibited monotropic nematic mesomorphism.  相似文献   

13.
1‐Methyl‐3,6,8‐trinitro‐2‐quinolone (1) behaved as the dienophile in Diels‐Alder reactions with dienes. When cyclopentadiene was used, cycloadduct 4 was obtained, which was then aromatized on treatment with triethylamine. In the reaction of 1 with hydrazone of 2‐butenal, phenanthridine derivative 7 was produced.  相似文献   

14.
The influence of the replacement of C=C bonds by isoelectronic B−N moieties on the reactivity of π-curved polycyclic aromatic hydrocarbons has been computationally explored by means of density functional theory calculations. To this end, we selected the Diels–Alder cycloaddition reactions of the parent corannulene and its BN-doped counterparts with either cyclopentadiene or maleic anhydride. In addition, the analogous reactions involving larger buckybowls, such as BN-hemifullerene, BN-circumtrindene, and BN-fullerene, have been also considered. It has been found that whereas corannulene behaves as a dienophile, its BN counterpart better acts as a diene. In contrast, the larger BN-curved systems cannot be used as dienes in Diels–Alder reactions, but undergo facile (i.e., low barrier) cycloaddition reactions with cyclopentadiene. The observed trends in reactivity, which cannot be directly explained by using typical frontier molecular orbital arguments, are quantitatively described in detail by means of state-of-the-art computational methods, namely the activation strain model of reactivity combined with the energy decomposition analysis method. The results of our calculations highlight the crucial role of the curvature of the system on the reactivity and its influence on the strength of the orbital interactions between the deformed reactants during their transformations.  相似文献   

15.
The aza‐Diels‐Alder cycloaddition reaction of cyclopentadiene with benzylideneanilines from methyl 12‐amino‐dehydroabietate is described via the Povarov reaction. © 2005 Wiley Periodicals, Inc. Heteroatom Chem 16:605–612, 2005; Published online in Wiley InterScience ( www.interscience.wiley.com ). DOI 10.1002/hc.20159  相似文献   

16.
The bulk free‐radical polymerization of 2‐[(N,N‐dialkylamino)methyl]‐1,3‐butadiene with methyl, ethyl, and n‐propyl substituents was studied. The monomers were synthesized via substitution reactions of 2‐bromomethyl‐1,3‐butadiene with the corresponding dialkylamines. For each monomer the effects of the polymerization initiator, initiator concentration, and reaction temperature on the final polymer structure, molecular weight, and glass‐transition temperature (Tg) were examined. Using 2,2′‐azobisisobutyronitrile as the initiator at 75 °C, the resulting polymers displayed a majority of 1,4 microstructures. As the temperature was increased to 100 and 125 °C using t‐butylperacetate and t‐butylhydroperoxide, the percentage of the 3,4 microstructure increased. Differential scanning calorimetry indicated that all of the Tg values were lower than room temperature. The Tg values were higher when the majority of the polymer structure was 1,4 and decreased as the percentage of the 3,4 microstructure increased. The Diels–Alder side products found in the polymer samples were characterized using NMR and gas chromatography‐mass spectrometry methods. The polymerization temperature and initiator concentration were identified as the key factors that influenced the Diels–Alder dimer yield. © 2000 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 38: 4070–4080, 2000  相似文献   

17.
Detailed density functional theory calculations definitively rationalize the preference for the endo cycloadduct (also known as endo rule) in text‐book thermal Diels–Alder reactions involving maleic anhydride and cyclopentadiene or butadiene. This selectivity is mainly caused by an unfavorable steric arrangement in the transition‐state region of the exo pathway which translates into a more destabilizing activation strain. In contrast with the widely accepted, orbital‐interaction‐based explanation for the endo rule, it is found that neither the orbital interactions nor the total interaction between the deformed reactants contributes to the endo selectivity. © 2013 Wiley Periodicals, Inc.  相似文献   

18.
Summary: Cyclopentadiene ( 1 ) was incorporated as a guest into the cavity of randomly methylated‐β‐cyclodextrin (me‐β‐CD) as a host, yielding the stable, water compatible cyclopentadiene/me‐β‐CD complex ( 1a ). We successfully attempted to use the synthesised complex in a Diels–Alder addition with a water‐soluble unsaturated polyester ( 2 ) derived from poly(ethylene glycol) and maleic anhydride. The reaction yielded a new type of polypseudorotaxane ( 3 ). Examination of the polypseudorotaxane ( 3 ) and a model inclusion complex of the starting unsaturated polyester with me‐β‐CD ( 2a ) showed that cyclodextrins are threaded onto the main chain in both cases. The cyclohexene moiety formed after the Diels–Alder addition does not act as a stopper, a dethreading process being evidenced and discussed.

The polypseudorotaxane synthesized here.  相似文献   


19.
Can achiral organocatalysts linked to chiral polyanionic metal oxide clusters provide good selectivity in enantioselective C?C bond formations? The answer to this question is investigated by developing a new active hybrid polyoxometalate‐based catalyst for asymmetric Diels–Alder reaction. Chirality transfer from the chiral anionic polyoxometalate to the covalently linked achiral imidazolidinone allows Diels–Alder cycloaddition products to be obtained with good yields and high enantioselectivities when using cyclopentadiene and acrylaldehydes as partners.  相似文献   

20.
Diels–Alder click reaction was successfully applied for the preparation of 3‐arm star polymers (A3) using furan protected maleimide end‐functionalized polymers and trianthracene functional linking agent (2) at reflux temperature of toluene for 48 h. Well‐defined furan protected maleimide end‐functionalized polymers, poly (ethylene glycol), poly(methyl methacrylate), and poly(tert‐butyl acrylate) were obtained by esterification or atom transfer radical polymerization. Obtained star polymers were characterized via NMR and GPC (refractive index and triple detector detection). Splitting of GPC traces of the resulting polymer mixture notably displayed that Diels–Alder click reaction was a versatile and a reliable route for the preparation of A3 star polymer. © 2007 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 46: 302–313, 2008  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