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.
The production of immobilized phospholipase A2 preparations by the covalent addition of the proteins of viper venom to Silokhroms containing epoxide groups is described. The dependence of the amount of protein bound to the support on its concentration and on the amount of epoxide groups in the sorbent has been investigated. The phospholipase activity of the preparations obtained did not depend on the amount of immobilized enzyme but was determined by diffusional factors.M. V. Lomonosov Institute of Fine Chemical Technology, Moscow. Translated from Khimiya Prirodnykh Soedinenii, No. 1, pp. 126–129, January–February, 1989.  相似文献   

2.
The preparation of partially saturated lightly functionalized styrene-butadiene block copolymers of polyA-block-polyB-block-polyA type (SBS) is described. The work involves epoxidizing partially hydrogenated SBS block copolymers using peracetic acid in a cyclohexane/water heterogeneous system. Five partially hydrogenated model polymers containing low levels of unsaturated aliphatic double bonds were used to study the epoxidation reaction and kinetics. The existence of the epoxide functional group on the product polymer was evidenced by IR and 1H-NMR spectra and the epoxide concentration was determined by direct titration. The partially hydrogenated SBS copolymers were more difficult to epoxidize than the unhydrogenated ones. The temperature dependence of the epoxidation rate was studied and the activation energy was determined as 8.8 kcal/mole of double bonds. © 1996 John Wiley & Sons, Inc.  相似文献   

3.
Three series of compounds based on the cyclohexene framework have been epoxidized by dimethyldioxirane. A pronounced dependence of epoxide diastereoselectivity on substituent has been observed. In addition there is a solvent influence on this stereoselectivity. The results have been explained by invoking steric, H-bonding, and dipole-dipole influences on the epoxide stereochemistry.  相似文献   

4.
Five-coordinate manganese(III) complexes of N, N'-bis(trifluoroacetylacetone)-1,2-ethylenediimine (tfacacen) have been synthesized and structurally characterized by X-ray crystallography. The presence of the electron-withdrawing -CF3 substituents enhances the electrophilicity of the metal center in these (tfacacen)MnX (X=Cl, N3, NCO, NCS) derivatives when compared with their (acacen)MnX (acacen=N, N'-bis(acetylacetone)-1,2-ethylenediimine) analogs. This is demonstrated by the increased propensity of the Mn(III) center in the tfacacen complexes to bind a sixth ligand. Binding studies were performed utilizing the upsilonN3 stretching frequency in (tfacacen)MnN3, which is sensitive to the coordination of a ligand at the vacant axial site. Of importance, cyclohexene oxide was shown to readily bind to (tfacacen)MnN3, thereby providing an opportunity for directly monitoring the dependence of the epoxide ring-opening process on the metal complex concentration. In this instance, as has been amply demonstrated in the (salen)CrX case, the ring opening of cyclohexene oxide was found to be second-order in [(tfacacen)MnN3], with an activation energy of 71.0+/-6.0 kJ/mol. In the presence of strongly coordinating anions or amine bases, the rate of epoxide ring opening by (tfacacen)MnN3 was greatly retarded. The manganese cyanate and thiocyanate complexes were examined in an effort to develop other initiators for epoxide ring opening which provide readily accessible infrared spectroscopic probes. Indeed, the thiocyanate ligand was found to be well-suited for monitoring the epoxide ring-opening reaction by infrared spectroscopy.  相似文献   

5.
The kinetic regularities of the change in the concentration of tert-butyl hypochlorite in the presence of the binary system (BS) styrene epoxide??p-toluenesulfonic acid in a tert-butanol solution were studied using iodometry and HPLC and compared with the data obtained earlier for hydroperoxide decomposition. The expressions for the rates of transformation of ButOCl, epoxide, and ROOH in the BS through the reactant concentrations are of the same type (the first order for the acid and the zero order for epoxide, ButOCl, and ROOH) and indicate that the reactions are related to epoxide heterolysis. Dioxygen ceases ROOH decomposition in the BS but exerts no effect on the decrease in the concentration of ButOCl, which efficiently inhibits the O2 uptake in the BS and almost an order of magnitude retards the accumulation of benzaldehyde (the product of styrene epoxide oxidation) with a low (up to 15%) decrease in the heterolysis rate. The inhibition effect is due to the heterolytic interaction of ButOCl with the carbocation formed by the cleavage of the three-membered ring of protonated styrene epoxide. The introduction of ButOCl in the BS decreases the stationary concentration of the carbocation and, as a consequence, the stationary concentration of phenylcarbene responsible for O2 uptake.  相似文献   

6.
蔡祖恽  王颖 《化学学报》1988,46(12):1191-1194
在不同浓度的三乙胺乙酸乙酯溶液中2-甲基环戊-1,3-二酮与氯代甲基乙烯酮(1b)反应可分别得三酮、α-氯代醇或环氧双酮. α-氯代醇经X射线单晶衍射分析结果推断出环氧双酮的环氧是β构型, 其异构体为α, 用相似的方法制得的同系环氧物也经X射线单晶衍射分析, 结果表明其环氧为β构型.  相似文献   

7.
The first catalytic method for the efficient conversion of epoxides to succinic anhydrides via one-pot double carbonylation is reported. This reaction occurs in two stages: first, the epoxide is carbonylated to a beta-lactone, and then the beta-lactone is subsequently carbonylated to a succinic anhydride. This reaction is made possible by the bimetallic catalyst [(ClTPP)Al(THF)2]+[Co(CO)4]- (1; ClTPP = meso-tetra(4-chlorophenyl)porphyrinato; THF = tetrahydrofuran), which is highly active and selective for both epoxide and lactone carbonylation, and by the identification of a solvent that facilitates both stages. The catalysis is compatible with substituted epoxides having aliphatic, aromatic, alkene, ether, ester, alcohol, nitrile, and amide functional groups. Disubstituted and enantiomerically pure anhydrides are synthesized from epoxides with excellent retention of stereochemical purity. The mechanism of epoxide double carbonylation with 1 was investigated by in situ IR spectroscopy, which reveals that the two carbonylation stages are sequential and non-overlapping, such that epoxide carbonylation goes to completion before any of the intermediate beta-lactone is consumed. The rates of both epoxide and lactone carbonylation are independent of carbon monoxide pressure and are first-order in the concentration of 1. The stages differ in that the rate of epoxide carbonylation is independent of substrate concentration and first-order in donor solvent, whereas the rate of lactone carbonylation is first-order in lactone and inversely dependent on the concentration of donor solvent. The opposite solvent effects and substrate order for these two stages are rationalized in terms of different resting states and rate-determining steps for each carbonylation reaction.  相似文献   

8.
贾涛  许建和  杨晟 《催化学报》2008,29(1):47-51
考察了多种载体对巨大芽孢杆菌ECU1001环氧水解酶的固定化.以大孔DEAE-纤维素离子交换树脂为载体时,固定化酶的活力回收达70%.进一步考察了温度和pH对固定化酶活力的影响,并使用该固定化酶进行了缩水甘油苯基醚对映选择性水解批次反应.结果表明,在较低的底物浓度下该固定化酶的稳定性较好,10批反应后仍然剩余72.4%的活力.  相似文献   

9.
10.
The temperature dependence of steady-shear viscosity and ionic conductivity were measured for a series of unreacted mixtures and partially cured, ungelled samples of diglycidyl ether of bisphenol-A (DGEBA) and an amine cross-linking agent, diamino diphenyl sulfone (DDS). Six stoichiometric ratios of epoxide groups to amine hydrogens were examined. Free volume expressions were used to model the temperature dependence of the conductivity and viscosity for the unreacted DGEBA-DDS mixtures. In addition, these expressions were combined to successfully correlate changes in viscosity and conductivity during the DGEBA-DDS polymerization prior to gelation. It also was demonstrated that the change in weight average molecular weight during polymerization could be interpreted from the dielectric data. Through studying variations in the stoichiometry, it was possible to examine the effects of changes in chemical structure and ion concentration on the fitted parameters in the free volume models. The inherent ion transport factor (ζ0) was found to be inversely proportional to the concentration of ions in the test samples. The fractional free volume for segmental motion (B) was found to increase with an increase in the glass transition temperature and to be a function of the rigidity of the polymer. ©1995 John Wiley & Sons, Inc.  相似文献   

11.
The addition reaction of styrene oxide (StO) with silk fibroin was studied in the presence of various salts in different solvents at 45–75°C. Some water was required to make StO react with silk padded with various salt solutions. The reaction rate increased with the salt concentration and reached a maximum value at a certain concentration of the salt. Padding with solutions of thiosulfate, cyanide, thiocyanate, bicarbonate, or carbonate resulted in high add-ons (to 65 mole/105 g) and low solubilities in HCl and NaOH aqueous solutions. The weight gains increased with the epoxide concentration and reached a constant value at a certain concentration of StO solution in ethanol, while they decreased slightly with epoxide concentration over 10% of StO solution in n-hexane. Histidine, lysine, arginine, tyrosine, and aspartic and glutamic acids were found to react. The reaction rate decreased with increasing solubility parameter of the solvent used, reached a minimum value about at 10 or at the solubility parameter of the epoxide, and then increased with the parameter. The StO–silk reaction may depend on the distribution of StO between aqueous salt and an organic solvent phases, and on the swelling of silk fiber in different aqueous salt solutions or in various organic solvents. The mechanism for this epoxide-silk reaction and the reactivity difference between StO and phenyl glycidyl ether toward silk fibroin are discussed in the light of the observed phenomena.  相似文献   

12.
Polybutadiene films were aged under air or high oxygen pressure (3.1 MPa). In both cases, the amount of epoxide formed was titrated. The results show that the epoxide formation rate is a decreasing function of oxygen concentration that validates the mechanism proposed by Mayo. According to this mechanism, epoxides are generated from the decomposition of β-peroxy alkyl radicals resulting from the addition of peroxy radicals on double bonds.  相似文献   

13.
The reaction of amino acids with functional groups in natural rubber has been studied by use of tritium labeled glycine as a tracer. The rate of reaction, in the latex phase, is found to be pH-dependent with optimum rate at about pH 8. Reaction is considered to occur with rubber epoxide groups, since the level of incorporation of glycine correlates closely with the epoxide groups concentration. Competitive reactions with other amino acids and the effect of dimedone on glycine incorporation are also reported.  相似文献   

14.
The first regiodivergent opening of unbiased epoxides (REO) providing the ring-opened products in high enantiomeric excess from racemic and exceptionally high enantiomeric excess from enantioenriched substrates in a double asymmetric process has been devised. It constitutes a more general case of the very important enantioselective openings of meso-epoxides. The dependence of the selectivity of ring opening on the epoxide’s substitution pattern was studied.  相似文献   

15.
Reaction of epoxy-diane resin with complex based on monoamide of phosphonic acid and ammonium chloride (MODAF) is studied. It is established that the MODAF equivalent in the reaction with epoxy resin is equal to 1/5 of its molecular weight. Epoxy resin is cured via the reactions of both monoamide and ammonium chloride. The dependence of the epoxide and hydroxyl groups has a stepped character over time, which is probably due to the existence of certain threshold concentrations of the curing coefficients. The investigated reaction is a new example of macroscopic coherence manifestation.  相似文献   

16.
The theory of branching processes is used to describe the polymer network formation resulting from the reaction of tetraepoxides with diamines using various initial compositions. Differences in reactivities of primary and secondary amine groups and the reaction between the epoxide groups and reaction-generated hydroxyl groups are taken into account; however, intramolecular reactions in the pregel stage are neglected. Expressions are derived for the critical epoxide conversion at the gel point, molecular weight in the pregel stage, changes in sol and gel fraction in the post-gel stage, and the concentration of elastically active network chains as a function of reaction conditions. The analysis of the simulation results shows that etherification reactions significantly raise the concentration of elastically active network chains of the mixture under stoichiometric excess of epoxide groups.  相似文献   

17.
Anion‐π catalysis functions by stabilizing anionic transition states on aromatic π surfaces, thus providing a new approach to molecular transformation. The delocalized nature of anion–π interactions suggests that they serve best in stabilizing long‐distance charge displacements. Aiming therefore for an anionic cascade reaction that is as charismatic as the steroid cyclization is for conventional cation‐π biocatalysis, reported here is the anion‐π‐catalyzed epoxide‐opening ether cyclizations of oligomers. Only on π‐acidic aromatic surfaces having a positive quadrupole moment, such as hexafluorobenzene to naphthalenediimides, do these polyether cascade cyclizations proceed with exceptionally high autocatalysis (rate enhancements kauto/kcat >104 m ?1). This distinctive characteristic adds complexity to reaction mechanisms (Goldilocks‐type substrate concentration dependence, entropy‐centered substrate destabilization) and opens intriguing perspectives for future developments.  相似文献   

18.
Bisketonate and alkoxide Ti(III) complexes derived from Zn reduction of Ti(IV) precursors were evaluated as catalysts for the living radical polymerization (LRP) of styrene initiated by Ti‐catalyzed epoxide radical ring opening and mediated by reversible termination with Ti(III). No polymerization occurred with tris(2,2,6, 6‐tetramethyl‐3,5‐heptanedionato)titanium (III), whereas dichlorobis(2,2,6,6‐tetramethyl‐3,5‐heptanedionato)titanium (IV) affords only a free radical polymerization. Preliminary living features were displayed by (iPrO)2TiCl2. Investigations of the effect of epoxide/Ti/Zn ratios, temperature, and nature of the epoxide demonstrated that (iPrO)3TiCl provides a linear dependence of Mn on conversion over a wide range of conditions with an optimum for [Sty]/[epoxide group]/[Ti]/[Zn] = 50/1/2/4 at 90 °C. However, the polydispersity could not be reduced below 1.4–1.5, with an initiator efficiency of 0.15. These results were rationalized in terms of a combination of decreased Ti oxophilicity and ligand exchange. The lowered oxophilicity decreases the initiation rate and broadens Mw/Mn. The fast alkoxide exchange promotes a weak dependence of the polymerization on reaction conditions and generates macromolecular Ti species with reduced ability to mediate LRP. Thus, while monofunctional epoxides provide homogeneous polymerizations and narrower Mw/Mn, difunctional initiators may lead to gel formation at high conversion. Nonetheless, all polymerizations were light gray to colorless and afforded white polymer. © 2005 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 43: 6028–6038, 2005  相似文献   

19.
The effects of the reducing agent, temperature, and epoxide/Ti and Ti/Zn ratios were investigated for the Cp2TiCl‐catalyzed living radical polymerization of styrene initiated by epoxide radical ring opening. No reduction of bis(cyclopentadienyl)titanium dichloride occurred with Cu, Devarda's alloy, Ni, Ce, Cr, Sn, Mo, and ascorbic acid, whereas Al, lithium nitride, Mn, Sm, and Fe led to free‐radical or poorly controlled polymerizations. The best results were obtained with Zn alloy, powder, or nanoparticles. Nano‐Zn provided the lowest polydispersity index values, highest initiator efficiency (IE), and fastest reaction rate while maintaining a well‐defined living polymerization. Progressively lower polydispersity was obtained with an increasing excess of Zn with an optimum at Cp2TiCl/Zn = 1/2. This was rationalized through the heterogeneous nature of Zn and its possible involvement in the reversible termination step. The polymerization was insensitive to light or dark conditions, and a linear dependence of Mn on the conversion was observed at all temperatures in the 60–130 °C range with an optimum at 70–90 °C. A stoichiometric 1/2 epoxide/Ti ratio provided low polydispersity (weight‐average molecular weight/number‐average molecular weight < 1.2) and high IE, whereas increasing the epoxide/Ti ratio to 1/3 maintained a low polydispersity index but decreased the IE. © 2006 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 44: 2156–2165, 2006  相似文献   

20.
Epoxidized glyceryl trioleate has been used as a model compound to extend previous studies on the mode of action of epoxides in the stabilization of poly(vinyl chloride) (PVC). Sheets of PVC containing 3% expoxidized glyceryl tri[1-14C]oleate were prepared on a hydraulic press and subsequently heated in an oven for various times to simulate heat processing. After separation of the polymer from the low molecular weight components by steric exclusion chromatography, there was only a negligible amount (max. 2%) of bonding to the polymer and progressive loss from 15 to 90% of epoxide function occurred with heat processing over the period from 0.2 to 40 min. The concentration of the chlorohydrin transformation product was estimated by high performance steric exclusion chromatography after formation of a suitable u.v. detectable derivative, but the levels found did not fully account for the overall loss of epoxide. To assess the potential of both epoxide and chlorohydrin for migration from plastic packaging into foods, sheets individually containing the epoxide and the corresponding chlorohydrin were placed in contact with a range of solvents, orange squash and edible oil. No migration was measurable at a limit of detection of 0.1 ppm (orange squash limit 0.5 ppm) after 10 days contact at 40, with the exception of edible oil where a level of 1.0 ppm was detectable.  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