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 318 毫秒
1.
用对苯二甲醛或间苯二甲醛与5,5-二甲基-1,3-环己二酮反应,在不同条件下得到了含有双4(H)-吡喃、1,4-二氢吡啶结构单元的双-氧杂蒽衍生物、双-吖啶衍生物、以及双-N-羟基吖啶衍生物.  相似文献   

2.
2-烯丙氧基-1,2-氧硼杂戊环(1)的侧链有较大化学活性。(1)可与醇类酯交换制得2-烷氧基-1,2-氧硼杂戊环,如2-正戊氧基-、2-环己氧基-、2-环辛氧基、2-氨基乙氧基-1,2-氧硼杂戊环;与乙二醇反应可制得2,2′-次乙二氧基-双-(1,2-氧硼杂戊环)(6),同样,也可制得1,2-丙二醇和2,3-丁二醇衍生物(7)(8)。(1)与三丁基硼和三烯丙基硼热交换可制得2-正丁基-1,2-氧硼杂戊环和2-烯丙基-1,2-氧硼杂戊环(10)。(10)的侧链活性更大,易与醇、胺类反应制得相应衍生物,如2-正丁氧基-和2-烯丙胺基-1,2-氧硼杂戊环,反应机理可能是烯丙基型重排。(1)与氯化亚砜作用可制得2-氯-1,2-氧硼杂戊环(13),另外还有二个破环、氯转位的副产物,3-氯丙基硼酸二烯丙酯及酐。(6)、(7)、(8)的双环相互作用可能形成网络状结构以及(13)的X-衍射结晶结构正在深入研究中。通过1,2-氧硼杂戊环类侧链的交换反应,合成侧链上具有碳、氮、氧、氯的衍生物,不仅表明侧链的多变性,而且表明环本身有相对稳定性。  相似文献   

3.
螺双二氢茚二酚(SPINOL)与N-溴代丁二酰亚胺经溴代反应制得6,6'-二溴-螺双二氢茚二酚(1);1与叔丁基二甲基硅基三氟甲磺酸(TBSOTf)经醚化反应得6,6'-二溴-7-羟基-7'-叔丁基二甲硅氧基-1,1'-螺二氢茚(2);SPINOL经羟基保护后与碘甲烷经双甲基化反应制得6,6'-二甲基-7,7'-双(1-甲氧甲氧基)-1,1'-螺二氢茚(4);4经脱保护后再与TBSOTf经醚化反应合成了6,6'-二甲基-7-羟基-7'-叔丁基二甲硅氧基-1,1'-螺二氢茚(6)。2和6为新化合物,其结构经1H NMR,13C NMR和HR-ESI-MS表征。  相似文献   

4.
采用微波辐射法合成双羟丙基封端聚硅氧烷,将响应面分析法应用于1,3-双(3-三甲硅氧基丙基)四甲基二硅氧烷的合成工艺,通过Box-Behnken中心组合实验,考察原料摩尔配比、反应时间、微波辐射温度和催化剂加量四个因素对1,3-双(3-三甲硅氧基丙基)四甲基二硅氧烷收率的影响,得到最佳微波合成工艺参数并建立多元回归数学模型。结果表明:微波合成最佳反应条件n(2-烯丙基三甲基硅氧烷)/n(1,1,3,3-四甲基二硅氧烷)=2.4∶1,反应时间103 min,微波温度95℃,催化剂加量19 mg/kg,得到产物收率可达94.46%。以优化条件下制备的1,3-双(3-三甲硅氧基丙基)四甲基二硅氧烷为原料,制备出了双羟丙基封端聚硅氧烷,目标产物转化率可达90.5%。采用IR、1HNMR、GC对产品结构进行了表征。  相似文献   

5.
吴世晖  武戈  陶凤岗  林子森 《化学学报》1987,45(11):1107-1111
2-(α-呋喃基)-2-苯基六甲基三硅烷在光照下可产生新型的有机硅活性中间体-α-呋喃基苯基硅烯.它与2,3-二甲基-1,3-丁二烯反应后,得到了相应的加成与插入反应产物;与环己烯反应后,再用甲醇分解所得硅杂环丙烷中间物得到了α-呋喃基环己基苯基甲氧基硅烷.  相似文献   

6.
钱长涛  王兵 《化学学报》1996,54(11):1084-1088
双(2-甲氧乙基环戊二烯基)氯化镧或氯化镱, 在四氢呋喃中, -20℃下,分别与苯乙炔基钠发生交换反应, 生成双(2-甲氧乙基环戊二烯基)苯乙炔基镧或镱, 产率分别为77%和66%。直接将无水三氯化镱或三氯化钇, 在四氢呋喃中, 冰水冷却下, 与两摩尔的双(三甲基甲硅烷基)氨基锂和2-甲氧乙基环戊二烯基钠进行一锅煮反应, 可得2-甲氧乙基环戊二烯基双(双三甲基甲硅烷基氨基)镱或钇, 产率分别为56%和72%。配合物经元素分析, 红外光谱, 核磁共振氢谱和质谱的鉴定, 它们可能是非溶剂化的, 含分子内配位键的中性单体配合物。  相似文献   

7.
2-(α-呋喃基)-2-苯基六甲基三硅烷在光照下可产生新型的有机硅活性中间体——α-呋喃基苯基硅烯.它与2,3-二甲基-1,3-丁二烯反应后,得到了相应的加成与插入反应产物;与环己烯反应后,再用甲醇分解所得硅杂环丙烷中间物得到了α-呋喃基环己基苯基甲氧基硅烷.  相似文献   

8.
双(2-甲氧乙基环戊二烯基)氯化镧或氯化镱,在四氢呋喃中,20℃下,分别与苯乙炔基钠发生交换反应,生成双(2-甲氧乙基环戊二烯基)苯乙炔基镧或镱.产率分别为77%和66%直接将无水三氯化镜或三氯化钇,在四氢呋喃中,冰水冷却下,与两摩尔的双(三甲基甲硅烷基)氨基锂和2-甲氧乙基环戊二烯基钠进行一锅煮反应,可得2-甲氧乙基环戊二烯基双(双三甲基甲硅烷基氨基)镱或钇,产率分别为56%和72%.配合物经元素分析,红外光谱,核磁共振氢谱和质谱的鉴定,它们可能是非溶剂化的,含分子内配位键的中性单体配合物.  相似文献   

9.
以新戊二醇、三氯化磷和乙醇为原料制得中间体 5 ,5 -二甲基 -1 ,3,2 -二氧磷杂环己烷 -2 -氧 ( DPDO) ,用DPDO与由芳醛和对苯二胺缩合制得的席夫碱 SB ~ SB 加成 ,合成了膨胀型阻燃剂 5 ,5 -二甲基 -2 -氧代 -2 -( N,N -对苯二胺基 -二苄基 ) -二 -1 ,3,2 -氧磷杂环己烷 ( DPPO ~ DPPO ) ,通过元素分析 ,IR,NMR和 MS对其结构进行表征 ,并对其在醇酸清漆中的应用进行了讨论  相似文献   

10.
以新戊二醇、三氯化磷和乙醇为原料制得中间体5,5-二甲基-1,3,2-二氧磷杂环己烷-2-氧(DPDO),用DPDO)与由芳醛和对苯二胺缩合制得的席夫碱SBⅠ~SBⅢ加成,合成了膨胀型阻燃剂5,5-二甲基-2-氧代-2-(N,N′-对苯二胺基-二苄基)-二-1,3,2-氧磷杂环己烷(DPPO Ⅰ~DPPOⅢ),通过元素分析,IR,NMR和MS对其结构进行表征,并对其在醇酸清漆中的应用进行了讨论.  相似文献   

11.
1,3-Bis(trichlorosilyl) cyclohexane was obtained by addition of HSiCl_3, to cyclohexadiene-1,3in the presence of H_2PtCl_6.6H_2O in isopropyl alcohol. The new compound was ethanolysed andmethylated to form di-methyl tetra-ethoxy disilyl, tri-methyl tri-ethoxy disilyl and terta-methyldi-ethoxy disilyl cyclohexanes. These di-to tetra-functional monomers were hydrolyzed by hydrochloric acid in ether. The di-functional monomer yielded cyclic dimer similar to octamethylcyclotetrasiloxane and the tri-functionalmonomer, a cyclic tetramer, while in the case of tetra-functional monomer a cyclic octamer wasobtained. These compounds have not been reported in literature.  相似文献   

12.
The copolymerization of tetraoxane with styrene catalyzed by BF3·O(C2H5)2 was studied at 30°C. to determine whether a cyclic monomer can copolymerize with a vinyl monomer. The formation of the copolymer was confirmed by elementary analysis of both benzene-soluble and benzene-insoluble fractions of the polymer obtained. It was found by gas chromatography that a fairly large amount of 4-phenyl-1,3-dioxane and a small amount of trioxane were formed in the present system, in addition to polymers. Roughly a third of the total amount of the monomers reacted was consumed in the formation of methanol-insoluble polymer, a third for 4-phenyl-1,3-dioxane, and another third for trioxane and unknown products which could not be indentified. The formation of these cyclic compounds during the copolymerization may be explained in terms of a back-biting (or intramolecular transacetalization) reaction. The cationic reactivity of tetraoxane was found to be similar to that of styrene on the basis of both the consumption rate of each monomer in the copolymerizing system and the composition of the methanol-insoluble polymer obtained.  相似文献   

13.
Two, functional, cyclic carbonate monomers, 5‐methyl‐5‐methoxycarbonyl‐1,3‐dioxan‐2‐one and 5‐methyl‐5‐ethoxy carbonyl‐1,3‐dioxan‐2‐one, were synthesized starting from 2,2‐bis(hydroxymethyl) propionic acid. The ring‐opening polymerization of the cyclic carbonate monomers in bulk with stannous 2‐ethylhexanoate as a catalyst under different conditions was examined. The results showed that the yield and molecular weight of polycarbonates were significantly influenced by the reaction conditions. The polycarbonates obtained were characterized by IR, 1H NMR, and differential scanning calorimetry. Their molecular weight was measured by gel permeation chromatography. The in vitro biodegradation and controlled drug‐release properties of the polycarbonates were also investigated. © 2003 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 41: 4001–4006, 2003  相似文献   

14.
The free radical polymerization of four methylated cyclic allylic sulfides was examined with reference to their polymerization volume shrinkage and the effect of ring size on reactivity. The compounds examined were 2‐methyl‐5‐methylene‐1,3‐dithiane ( 5 ) (solid), 2‐methyl‐6‐methylene‐1,4‐dithiepane ( 6 ) (liquid), 6‐methyl‐3‐methylene‐1,5‐dithiacyclooctane ( 7 ) (liquid), and 6,8‐dimethyl‐3‐methylene‐1,5‐dithiacyclooctane ( 8 ) (liquid). The monomers were stable materials not requiring any special handling or storage conditions. They were polymerized in bulk using thermal azobisisobutyronitrile (AIBN, VAZO88) and photochemical initiators (Ciba DAROCUR 1173) and in benzene solutions (AIBN, 70 °C). The six‐membered ring monomer 5 was unreactive whereas seven‐membered ring monomer 6 polymerized to high conversion in bulk. In addition, 6 did not polymerize in benzene solution at 70 °C at [ 6 ] = 1.25M. Eight‐membered ring monomers 7 and 8 polymerized in bulk to complete conversion with thermal and photochemical initiators to give lightly crosslinked materials. Near complete conversion to soluble polymers could be obtained in solution polymerizations in benzene. Soluble polymers were also obtained in photochemical initiated bulk polymerizations by lowering initiator concentrations or length of irradiation. The methyl substituent had no effect on which allylic carbon–sulfur bond fragmented in the ring‐opening step. The polymerization volume shrinkages of monomers 7 and 8 were 1.5 and 2.4% respectively and together with monomer 4 (1.5–2.0% shrinkage) are the best available liquid free radical ring‐opening monomers that can be polymerized in bulk at room temperature. © 2000 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 39: 202–215, 2001  相似文献   

15.
It was determined whether trioxane, a cyclic formal, can copolymerize with styrene, a vinyl monomer, in the presence of BF3·O(C2H5)2 catalyst at 30°C. The methanol-in-soluble fraction after extraction with benzene was found to contain the copolymer of styrene and trioxane, thus demonstrating that trioxane can copolymerize with styrene In this case the amount of the methanol-insoluble polymer was less than that of the total monomer consumed, as determined by gas chromatography. This was found to be caused partly by the formation of the cyclic oligomer, 4-phenyl-1,3-dioxane. The relative reactivity of styrene was qualitatively found to be larger than that of trioxane, not only from the rate of monomer consumption but also from the composition of the methanol-insoluble polymer obtained. In a nonpolar solvent the reactivity of trioxane increased, and the difference in reactivity between the two monomers decreased. Indeed, an apparent monomer reactivity ratio might be obtained from the relationship between the monomer composition and the monomer consumption rate or the composition of the methanol-insoluble polymer, but it did not have a quantitative meaning because of the complexity of the copolymerization reaction.  相似文献   

16.
A new cyclic monomer, 2-oxo-3-methylene-5, 6-diphenyl-1, 4-dioxan, was synthesized. Thestructure of the intermediates and the monomer were determined by IR,~1H NMR,~(13)C NMR andelemental analysis. This new monomer is different from other cyclic monomers in this series,it isa solid (mp 108--109℃)and not very reactive, but still can undergo free radical ring-openingpolymerization. The free radical polymerization was carried out at 130℃. The structure of theresulting polymer was discussed and charaterized by IR, ~1H NMR, ~(13)C NMR and elementalanalysis. The molecular weight of the polymer was estimated by viscosity determination.  相似文献   

17.
A new cyclic carbonate, 2,2-ethylenedioxypropane-l,3-diol carbonate (EOPDC), was synthesized through a two-step reaction from dihydroxyacetone dimer, and polymerized in bulk initiated by Sn(Oct)2 to give a high molecular weight polycarbonate. The structure of monomer and the polymer were characterized by FT-1R, ^1H NMR, ^13C NMR. The cytotoxicity of the obtained polycarbonate was investigated by MTT assay.  相似文献   

18.
Novel monofunctional brominated benzoxazine 3‐(2,4,6‐tribromophenyl)‐3,4‐dihydro‐2H‐1,3‐benzoxazine (P‐bra) and bifunctional brominated benzoxazine 6,6′‐bis(3‐(2,4,6‐tribromophenyl)‐3,4‐dihydro‐2H‐1,3‐benzoxazinyl) isopropane (B‐bra) were prepared and highly thermally stable polybenzoxazines were obtained by the thermal cure of the corresponding benzoxazines monomers. The chemical structures of these novel monomers were confirmed by FITR, 1H‐NMR and elemental analysis. FTIR spectra and differential scanning calorimetry (DSC) suggested that the polymerization was thermally initiated and occurred via ring‐opening of the monomer in each case. Thermogravimetric analysis (TGA) indicated that brominatation could have a profound effect on increasing char yield and on thermal degradation temperatures. The results of UL‐94 burn test showed that the polybenzoxazines prepared from P‐bra and B‐bra had good flame retardance. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

19.
The free‐radical cyclopolymerization of diallyl ether (1) and methyl α‐(allyloxymethyl)acrylate (2) has been modeled with the B3LYP/6‐31G* methodology, by making use of model compounds for the growing radicals. The cyclization of both monomers is exo, with activation barriers of 5.33 and 9.82 kcal/mol, respectively. To account for the polymerizabilities of these monomers, competing reactions have also been modeled. Although both monomers have a lower barrier for homopolymerization than for cyclization, cyclization dominates due to entropy. This explains the high cyclopolymerization vs. homopolymerization of monomer 2, although its monofunctional counterpart has been reported to homopolymerize well. It has also been shown that the degradative chain transfer by H‐abstraction from the allylic carbon is not effective with this monomer. Poor cyclopolymerization of the monomer 1 has been demonstrated by modeling the degradative chain transfer by H‐abstraction from the allylic carbon, which has been shown to compete very efficiently with polymerization reactions. Additionally, intermolecular propagation reaction has been shown to be facile due to cyclization, since the attacking monomer adopts a cyclic structure. © 2006 Wiley Periodicals, Inc. Int J Quantum Chem, 2007  相似文献   

20.
This work deals with the cationic ring‐opening polymerization of the cyclic thiocarbonates 5‐benzoyloxymethyl‐5‐methyl‐1,3‐dioxane‐2‐thione ( 1 ), 5,5‐dimethyl‐1,3‐dioxane‐2‐thione ( 2 ), and 4‐benzoyloxymethyl‐1,3‐dioxane‐2‐thione ( 3 ). The polymerization was carried out with 2 mol % trifluoromethanesulfonic acid, methyl trifluoromethanesulfonate, boron trifluoride etherate, or triethyloxonium tetrafluoroborate as the initiator to afford the polythiocarbonate with a narrow molecular weight distribution accompanying isomerization of the thiocarbonate group. The molecular weight of the obtained polymer could be controlled by the feed ratio of the monomer to the initiator and increased when the second monomer was added to the polymerization mixture after the quantitative consumption of the monomer in the first stage. The block copolymerization of 2 and 3 was also achieved, and this supported the idea that the cationic ring‐opening polymerization of these monomers proceeded via a living process. The order of the polymerization rate was 3 > 2 > 1 . The cationic ring‐opening polymerization of 1 and 3 involved the neighboring group participation of ester groups according to the polymerization rate and molecular orbital calculations with the ab initio method. © 2002 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 41: 185–195, 2003  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