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.
A formal synthesis of a powerful cholesterol inhibitor, ezetymibe 1, is described. The crucial step of the synthesis is based on Cu(I)-mediated Kinugasa cycloaddition/rearrangement cascade reaction between terminal acetylene derived from acetonide of L-glyceraldehyde and suitable C,N-diarylnitrone. The adduct with (3R,4S) configuration at the azetidinone ring, obtained with high stereoselectivity, was subsequently subjected to deprotection of the diol side chain followed by glycolic cleavage and base-induced isomerization at the C3 carbon atom to afford the (3S,4S) aldehyde, which has been already transformed into ezetimibe by the Schering-Plough group.  相似文献   

2.
Solid-state 13C NMR spectroscopy has been used to determine the decrease in C?C bonds, formation of crosslinks and cis to trans isomerization during the γ irradiation of (a) > 99% cis, 1,4-polybutadiene, (b) 54% trans, 41% cis, 1,4-polybutadiene, and (c) 86% 1,2-polybutadiene. G(-cis C?C) and G(-trans C?C), were similar and decreased with dose from ≈ 40 for 0-1 MGy to 5 for 5-10 MGy. G(-double bonds) and G(crosslink) were comparable, indicating that crosslinking occurred through the double bonds. G(crosslink) was much higher than values derived from physical properties, confirming that NMR measures the total of inter- and intramolecular crosslinking (cyclization). The 1,2 polybutadiene was much more sensitive to crosslinking, and a value of G(-C?C) = 240 was obtained at low doses. Crosslinking evidently proceeds by a kinetic chain reaction in all three types of polybutadiene.  相似文献   

3.
采用密度泛函理论, 在B3LYP/6-31G**水平上, 研究了气相和水溶剂中, α-羟基化吡咯烷亚硝胺(α-hydroxylation-NPYR, A)代谢为终致癌物重氮氢氧化物(B)、重氮烷阳离子(C)和氧离子(D), 以及C与鸟嘌呤碱基相互作用的反应机理. 化合物A代谢为终致癌物, 涉及异构化和质子化过程, 是相对容易进行的放热反应. 终致癌物C与鸟嘌呤在N7位形成DNA加合物F和G的反应, 遵循SN2机理. 加合物G由F异构形成, 且有相对高的异构化能(气相: 244.77 kJ/mol; 水溶剂中: 234.83 kJ/mol), 这与实验上得到加合物G是主要癌变物的结果一致.  相似文献   

4.
Hydroalumination of aryldialkynylphosphines RP(C≡C-(t)Bu)(2) (R = Ph, Mes) with equimolar quantities of diethylaluminum hydride afforded mixed alkenyl-alkynyl cyclic dimers in which the dative aluminum-phosphorus bonds are geminal to the exocyclic alkenyl groups. Addition of triethylaluminum to isolated 1 (R = Ph) or to the in situ generated species (R = Mes) caused diethylaluminum ethynide elimination to yield the arylethylphosphorus dimers 2 and 3. These possess a chair-like Al(2)C(2)P(2) heterocycle with intermolecular Al-P interactions. The boat conformation (4) was obtained by the reaction of (t)Bu-P(C≡C-(t)Bu)(2) with di(tert-butyl)aluminum hydride. Despite being dimeric, 2 behaves as a frustrated Lewis pair and activates small molecules. The reaction with carbon dioxide gave cis/trans isomeric AlPC(2)O heterocycles that differ only by the configuration of the exocyclic alkenyl unit. Four isomers resulted from the reaction with phenyl isocyanate. This is caused by cis/trans isomerization of the initial C=O adduct and subsequent rearrangement to the AlPC(2)N heterocycle, being the C=N adduct.  相似文献   

5.
Low molecular weight polymers and copolymers of butadiene were grafted with styrene. The graft products were then crosslinked by using dicumyl peroxide as initiator. The optimum peroxide concentration was established (5 phr). Infrared analysis showed that the reactivity of 1,2-vinyl and that of 1,4-trans double bonds in styrene-grafted polybutadiene is similar. Crosslinking of the graft product seems to involve a radical-chain polymerization of double bonds in the polymer. The reaction rate is proportional to the square root of peroxide concentration and to the concentration of polymer double bonds. Activation energy, reaction heat, reaction order, and crosslinking efficiency were also determined from DSC measurements. No relation was found between the activation energy of crosslinking and the molecular weight of backbone polymer or density of grafting. Crosslinking efficiency was to 25–50 crosslinks per molecule of decomposed peroxide. The crosslinking efficiency for grafted butadiene–styrene copolymers is somewhat lower than that for grafted polybutadienes. From thermogravimetric measurements it was found that the crosslinked grafted polymers show lower resistance to thermal degradation than ungrafted polymers.  相似文献   

6.
The low-temperature chemical crosslinking of a copolymer of vinylidene fluoride and tetrafluoroethylene was studied through various physicochemical methods. The reaction was conducted in solution in the presence of diethylenetriamine as a crosslinking agent. The penetration of the mixture molecules and the crosslinking agent only in the amorphous phase of the copolymer was provided via selection of the ratio between a good solvent (dimethylformamide) and a poor solvent (ethanol). Owing to this, the crosslinking reaction occurred in the amorphous phase and hardly involved the crystals. This outcome was confirmed by wide-angle X-ray diffraction data. The structural and chemical changes in the amorphous phase during crosslinking were recorded with the use of IR spectroscopy and differential scanning calorimetry. It is shown that crosslinking is accompanied by the formation of C=C bonds in the copolymer chains. The study of high-voltage polarization and conductivity during exposure to bipolar rectangular pulses suggested that crosslinking leads to an increase in the carrier concentration. It was found that the surface potentials in the films increase with an increase in the number of high-voltage pulses applied to the sample. This circumstance is attributed to the fact that the double bonds formed in the copolymer chains can effectively trap negatively charged carriers.  相似文献   

7.
Infrared spectroscopy and differential scanning calorimetry were used to study the imidization and crosslinking kinetics of norbornenyl-capped addition type of polyimide resins (designated PMR for polymerization of monomer reactants). The spectral and thermal analyses were performed on resin specimens that had been isothermally aged at temperatures appropriate for imidization (120–204°C) and crosslinking (275–325°C). Imidization occurs rapidly (about 10?2 min?1) at short times, but, at times longer than about 0.5 hr, the rate decreases significantly (about 10?4 min?1). The crosslinking reaction exhibits first-order kinetics during the initial part of the reaction, and its rate appears to be limited by the reversion of the norbornenyl Diels-Alder adduct. Both the first-order dependence and the activation energy (about 44 kcal) are consistent with this interpretation. The total heat evolved per mole of end cap during crosslinking shows an inverse dependence on the molecular weight of the imido prepolymers. This reflects the effect of end-cap dilution and decreased mobility of the larger oligomers.  相似文献   

8.
The C + PH(3) reaction is one of the simplest gas-phase processes which can produce molecular species containing P-C bonds. It could be of astrophysical importance and a reference for other phosphine reactions with carbon-containing molecular radicals. The dynamical aspects have been studied theoretically by quasi-classical trajectory methods in order to determine its rate as a function of the temperature, the branching ratios, and the molecular mechanisms. We have obtained a T(0.2) dependence of the capture rate. The total rate is affected by the existence of relatively high-lying saddle points for the isomerization of the CPH(3) complex but get a value of 0.82·10(-10) cm(3) s(-1) at 300 K, which is considered quite high for a neutral-neutral reaction and higher than those of similar reactions. Moreover, the total rate presents a weak dependence with the temperature. Our results indicate that several products containing P-C bonds are formed, the main reaction channel being the generation of HPCH + H.  相似文献   

9.
A novel crosslinker for thermally reversible covalent (TRC) linking of halide-containing polymers is suggested. Chlorine-containing polymers such as chloromethylstyrene copolymers, chlorinated polypropylene, polyvinylchloride, chlorinated polyisoprene, and polyepichlorohydrin were crosslinked with potassium dicyclopentadienedicarboxylate (KDCPDCA). The crosslinker was prepared by reacting potassium ethoxide with dicyclopentadienedicarboxylic acid. Because of the low solubility of KDCPDCA in organic solvents, a phase transfer catalyst, benzyltrimethyl-ammonium bromide, was employed for the crosslinking reaction. The crosslinking reaction occurred at a higher rate in a polar solvent, such as dimethylformamide, than in a nonpolar one, such as toluene, and was affected by the nature of the chlorine-containing polymer. Some of the polymers crosslinked even at room temperature. The chain-extending reaction between KDCPDCA and a α,ω-dihalide compound such as α,α′-dichloro-p-xylene, 1,4-dichlorobutane, or 1,4-dibromobutane also was carried out to obtain linear oligomers. The IR spectra indicated that the crosslinking and chain-extending reactions were based on the esterification between the halide carbon bonds of the polymer and the COOK groups of KDCPDCA. The flowability at 195 °C and solubility on heating in a dichlorobenzen-maleic compound mixture of the crosslinked polymers indicated that the TRC crosslinking occurred via the reversible Diels–Alder cyclopentadiene/dicyclopentadiene conversion as long as the polymer was thermally stable and did not contain olefinic CC bonds. The TRC linking also was confirmed by the rapid decrease of the specific viscosity of the obtained linear oligomers on heating. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 4390–4401, 1999  相似文献   

10.
11.
The [4 + 2] cycloaddition of trans,trans-2,4-hexadiene with C(60) proceeds via a concerted mechanism with retention of stereochemistry in the cycloadduct 1a. However, when cis,trans-2,4-hexadiene reacts with C(60), isomerization of the cis,trans to the thermodynamically more stable trans,trans isomer occurs. Subsequently, the cis,trans diene isomerized to the trans,trans isomer and cycloadds to C(60), to form adduct 1a. When the reaction is carried out at higher temperatures, the formation of cycloadduct 1b is also obtained. This result is consistent with a concerted cycloaddition of cis,trans-2,4-hexadiene with C(60), which is more reactive at elevated temperatures and leads to the formation of the Diels-Alder adduct 1b.  相似文献   

12.
The kinetics and thermodynamics of the liquid phase isomerization of an α- and β-pinene mixture on a Pd/C catalyst were studied. The effects of pinene concentration, catalyst particle size, stirring rate, reaction temperature (293–423 K) and hydrogen pressure (0.5–11 bar) on the rate of α- and β-pinene isomerization were investigated. The reaction rate of the α-pinene isomerization has the first order with respect to the α-pinene concentration and 0.5 order with respect to the hydrogen pressure. The thermodynamic parameters of the isomerization (Gibbs energy, reaction enthalpy and reaction entropy) and the equilibrium ratio of pinenes under the conditions studied were determined. The general scheme of the reaction mechanism of α- and β-pinene isomerization over the Pd/C catalyst was proposed.  相似文献   

13.
Currently, efforts are being made to reduce xenograft calcification and to optimize biomechanical properties by applying different crosslinking methods and techniques. Porcine aortic leaflets could be stabilized with a bisepoxy compound, butane‐1,4‐diol diglycidyl ether (BDDGE), under acidic (pH 4.5) or alkaline conditions (pH > 8.5). Maximum values of the shrinkage temperature (83°C), which are comparable to glutaraldehyde‐crosslinked materials, are achieved at pH 9.0 and 10.0 within 48 h with a BDDGE concentration of 4 wt6percnt;. The crosslinking efficacy decreased at higher pH values or at prolonged reaction times due to a higher proportion of one‐side or masking reactions. Crosslinking of aortic leaflets at pH 4.5, which consisted in a reaction between the epoxide groups of BDDGE and the carboxylic acid groups of the tissue, afforded crosslinked material with a shrinkage temperature of 76°C after 7 days of reaction.  相似文献   

14.
The effect of α‐methyl styrene dimer (AMSD), which is used as a scorch retarder, on the reaction mechanisms of the chemical crosslinking of polyethylene (PE) with dicumyl peroxide (DCP) at high temperatures was investigated using electron spin resonance. When AMSD was added to PE containing DCP, the AMSD radical was observed; however, the PE alkyl radical or allyl radical presence was not detected. At 145 °C, crosslinking was obstructed as a result of the reaction between AMSD and alkyl radicals. As the temperature increased, AMSD fragmented to form 2‐phenyl‐2‐propyl and double bonds in PE. This generation of double bonds, however, accelerated crosslinking at 180 °C and was more effective than when AMSD was not present. © 2001 John Wiley & Sons, Inc. J Polym Sci Part A: Polym Chem 39: 2151–2156, 2001  相似文献   

15.
Increasing evidence indicates that sulfur-containing molecules can play important roles in the activity of platinum anticancer drugs. Although nuclear DNA is retained to be the ultimate target, these platinum compounds can readily react with a variety of other substrates containing a soft donor atom, such as proteins, peptides, and low molecular weight biomolecules, before reaching DNA. In a recent study it was demonstrated that the DNA platination rate of a trans-geometry antitumor drug was dramatically enhanced by methionine binding, thus suggesting that the thioether could serve as a catalyst for DNA platination. In this work we performed detailed studies on the reactions of a widely investigated and very promising trans-platinum complex having two iminoethers and two chlorido ligands, trans-EE, with methionine (Met) and guanosine 5'-monophosphate (GMP). The results show that in the reaction of trans-EE with methionine the bisadduct is the dominant species in the early stage of the reaction. The reaction is also influenced by chloride concentration: at low NaCl the bis-methionine adduct is formed in preference, whereas the monoadduct is favored at high NaCl concentration. Not only the monomethionine complex, trans-PtCl(E-iminoether)(2)(AcMet), but also the bis-methionine adduct, trans-Pt(E-iminoether)(2)(AcMet)(2), which has already lost both leaving chlorides, can react with GMP to form the ternary platinum complex trans-Pt(E-iminoether)(2)(AcMet)(GMP). The latter reaction discloses the possibility of direct coordination to DNA of a platinum-protein adduct, in which the two carrier ligands remain intact; this is not the case of cis-oriented platinum complexes, like cisplatin, for which formation of a ternary complex is usually accompanied by loss of at least one carrier ligand. Interestingly, isomerization from S to N coordination of one methionine takes place in the bis-methionine complex at neutral pH, while the monoadduct appears to be stable. The shift from S to N coordination of one methionine in the trans-bis-methionine adduct can easily account for the obtainment of the cis isomer in the bis-chelated Pt(Met-S,N)(2) end product.  相似文献   

16.
通过试管倒立法和流变学测试对交联型聚酰胺胺凝胶的形成进行了判定。实验结果表明,直观的试管倒立法对聚酰胺胺凝胶形成的判定是有效的。通过对影响凝胶形成的各因素分析发现,在R=1.25、反应温度50℃、甲醇溶剂中制备的凝胶在单体质量浓度大于22%时能够形成凝胶,且符合IPN凝胶的制备要求。通过对聚酰胺胺(PAMAM)、N,N-亚甲基双丙烯酰胺(MBA)和N-氨乙基哌嗪(AEPZ)的红外分析可以看出,本文中MBA与AEPZ可以发生交联反应,生成PAMAM,同时对加入LiBr的PAMAM和未加入LiBr的PAMAM红外分析表明,交联聚酰胺胺中含有丰富的氢键相互作用,为材料的自修复提供了可能。通过对PAMAM-1(单体质量浓度22%)、PAMAM-2(单体质量浓度27%)和PAMAM-3(单体质量浓度32%)的平衡溶胀率测试可以看出,PAMAM-1具有最大的溶胀率。通过对三者的动态模量测试,PAMAM-1、PAMAM-2和PAMAM-3的G′分别为5000Pa、15000Pa和24000Pa,说明凝胶具有良好的弹性。以PAMAM-3为例表征了材料的自修复过程,表明制备的交联型聚酰胺胺能够通过氢键及物理的相互作用对材料进行修复,在药物缓释等方面有潜在应用。  相似文献   

17.
The kinetic studies of the crosslinking reaction of a nadic end-capped imide model compound, N,N′-(oxydi-3,4′-phenylene) bis(5-norbornene-2,3-dicarboximide), a bisnadimide, in thermal and microwave processes were investigated. The conversion of the endo isomer to exo isomer proceeds at a much lower temperature than the crosslinking reaction. The crosslinking reaction was monitored by the combined decrease in the infrared absorptions of the endo and exo isomers at 840 and 780 cm−1, respectively. The decrease in the concentration of starting materials follows first-order kinetics in the thermal and microwave processes. At the same temperatures (230 or 280°C), the crosslinking reaction proceeds at about 10 times faster in the microwave process than in the thermal process. Solid-state 13C-NMR showed no significant loss in C=C double bond resonance in the cured products by comparison with the starting material. This study provides direct evidence that the microwave process may be an efficient method to cure nadic end-capped polyimides. © 1998 John Wiley & Sons, Inc. J. Polym. Sci. A Polym. Chem. 36: 2653–2665, 1998  相似文献   

18.
Thermal cis-trans isomerization and decomposition of polyacetylene film prepared with a Ti(OC4H9)4–Al(C2H5)3 (Al/Ti = 4) system were investigated under inert gas or in vacuum by means of thermal analysis and infrared spectroscopy. Thermograms of differential thermal analysis of cis-polyacetylene revealed the existence of two exothermic peaks at 145 and 325°C and one endothermic peak at 420°C which were assigned to cis-trans isomerization, hydrogen migration accompanied with crosslinking reaction, and thermal decomposition, respectively. The isomerization was followed by infrared spectroscopy over the temperature range 75–115°C. The reaction did not obey simple kinetics. The apparent activation energy for the cis-trans isomerization was 17.0 kcal/mole for the polymer containing 88% cis configuration and increased with increasing trans content up to 38.8 kcal/mole for 80% trans content.  相似文献   

19.
When reacting ethylene-vinyl acetate copolymers (EVA) with sodium alcoholates in alcohol, the alcoholysis of EVA is accompanied by crosslinking reactions. Crosslinking can be detected by the increase in torque in a plastograph. This paper reports on investigations on the mechanism of the crosslinking reaction. Both crosslinking via C?C bonds and formation of a thermally reversible network via polymer alcoholate structures can be found. If an ethylene-vinyl alcohol copolymer is obtained by complete saponification of EVA, however, only thermally reversible crosslinking can be demonstrated. This leads to the conclusion that irreversible crosslinking is established through a reaction of the acetate side chains. The reaction products of the low-molecular weight model substances, pentanol-3-acetate and sodium isopropylate in isopropanol, are analyzed by means of gas chromatography. These results suggest a partial Claisen condensation as a mechanism of the irreversible crosslinking process.  相似文献   

20.
A convenient zinc‐promoted [4+3] cycloaddition of a carbonyl ene–yne with simple dienes was first achieved. This reaction provided an efficient strategy to prepare various cyclohepta[b]furan rings by cascade cycloadditions. Additionally, a multicomponent reaction of dione, alkynal, and diene was also reported, which exhibited a novel strategy for selective creations of C?O bonds and C?C bonds.  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