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 905 毫秒
1.
以5,6-二溴-2,3-二氰基氢醌为底物,在pH4.50条件下,用分光光度法考察了FeSO4对漆树漆酶催化氧化反应的影响.结果表明,在本文实验条件下SO2-4离子对酶促反应的影响基本可以忽略,而Fe2+离子对漆酶的催化氧化反应则有明显的竞争性抑制作用.研究证实,Fe2+离子的抑制作用是通过它还原酶促反应的产物半醌自由基阴离子来实现的.  相似文献   

2.
醌环的取代基对醌类化合物进行光合作用的电子传递过程有很大影响. 我们设想若能在保持与醌环上的甲基相近体积的条件下, 将其中一个甲基换成氯原子, 这样, 这类化合物醌环上的电子密度显然会低于质体醌, 从而进一步研究它们在光合作用中的作用, 有助于进一步阐明光合作用机制, 以3-氯-2-甲基苯胺为原料, 经氧化, 自由烷基化反应, 合成了十六个具有不同长侧链结构的质体醌类似物.  相似文献   

3.
金属锡作用下烯丙基溴与α,β-不饱和醛区域选择性加成,只生成1,2-加成产物.烯丙基卤代物和锡与1,4-醌类反应,不经分离,直接氧化可得到烯丙基取代醌类.  相似文献   

4.
周金林  邱雪芬  季凤英  马林  黄志纾  古练权 《有机化学》2004,24(7):819-821,J005
以1,4-对醌、芳肼为原料,通过Aza-Nenitzescu反应、氧化银氧化得到吲唑醌,亲核试剂(各种芳香胺)在Ce3 催化下与吲唑醌发生迈克尔加成反应,得到一系列5-芳氨基基取代的吲唑醌.化合物结构经元素分析、IR及1H NMR证实.初步的生理活性研究表明,这些化合物具有良好的抗哮喘活性.  相似文献   

5.
3,5,3',5'-四叔丁基-苄烯-4,4'-醌的合成和分子结构   总被引:2,自引:0,他引:2  
酚与铜的相互作用的研究有助于阐明生物体中酶催化氧化的作用过程和有机合成的催化机理。文献中对于2,4-或2,6-二取代酚的氧化偶联反应已有所报道本文报道了2,6-二叔丁基4-甲基酚钠在无水无氧条件和无水氯化亚铜或氯化铜作用下,可发生高效氧化偶联反应,且偶联反应发生在对位甲基位置上,生成3,5,3’,5’-四叔丁基-苄烯4,4’-醌。本文还表征了该化合物的分子结构。  相似文献   

6.
高翔  黄邦州  吴世晖 《化学学报》1991,49(8):827-832
金属锡作用下烯丙基溴与α,β-不饱和醛区域选择性加成,只生成1,2-加成产物。烯丙基卤代物和锡与1,4-醌类反应,不经分离,直接氧化可得到烯丙基取代醌类。  相似文献   

7.
采用电化学循环伏安法(CV)和原位拉曼光谱(in situ Raman)对窄能隙共轭高分子聚{吡咯 2,5 二[(对二甲氨基)苯甲烯]}(PPDMABE)的电化学行为进行了研究.结果表明,在不同pH值NaNO3溶液中, PPDMABE的电化学氧化还原过程中存在吡咯环的氧化态结构与芳式和醌式结构之间的转变.聚合物在氧化态时吡咯环主要以氧化态存在,而还原态以芳式和醌式结构吡咯环为主.PPDMABE在酸性和中性介质中氧化态吡咯环以质子化的状态存在,而在碱性溶液中氧化态吡咯环既有质子化态,又有去质子化态的.在酸性条件下,PPDMABE较易发生氧化还原反应,而在碱性条件还原反应则较难发生.  相似文献   

8.
采用密度泛函理论方法在M06-2X/6-311G*水平上模拟了不同反应条件下, TiO2对苯甲醛的光催化还原和氧化的反应. 计算结果表明, 苯甲醛的光催化还原和氧化反应均可在常温下发生; 在缺氧但有乙醇存在的条件下, 乙醇分子可与氧化性物质发生反应, 生成醇自由基, 苯甲醛主要发生光催化还原反应生成苯甲醇; 在有氧气但无乙醇存在条件下, 还原性的光生电子被氧气捕获, 避免了苯甲醛被还原, 主要发生光催化氧化反应生成苯甲酸.  相似文献   

9.
朱本占  张静  唐苗  黄春华  邵杰 《化学进展》2022,34(1):227-236
卤代醌是一类卤代芳烃类环境污染物的致癌中间体,也是在饮用水中新发现的氯化消毒副产物。我们最近发现卤代醌和 H2O2 或有机氢过氧化物体系可以不依赖过渡金属离子,而产生高活性的羟基/烷氧自由基和醌氧/醌碳自由基。目前尚不清楚这些卤代醌类致癌物和氢过氧化物共存能否诱导 DNA 产生氧化损伤和修饰,以及其潜在的分子机制是什么。我们的研究发现 DNA 在四氯-1,4-苯醌/H2O2体系中可被氧化产生 8-氧脱氧鸟苷、DNA 链断裂和三种甲基氧化产物,这些反应不依赖过渡金属离子,且由于卤代醌与 DNA 的嵌入作用而导致其氧化作用增强。其他卤代醌也观察到了类似的现象,而且通常比经典的 Fenton 体系更有效。我们进一步将研究从纯化的 DNA 扩展到了活细胞的基因组 DNA。同时还发现卤代醌和有机氢过氧化物(如叔丁基过氧化氢或在正常生理条件下产生的 13S-过氧羟基-9Z,11E-十八碳二烯酸(13-HPODE))共存时,可通过独特的醌氧自由基介导机制诱导 DNA 氧化生成致突变性更强的咪唑啉酮类产物 dIz。这些发现为解释普遍存在的卤代醌类致癌中间体和消毒副产物的潜在基因毒性、致突变性和致癌性提供了新思路。  相似文献   

10.
冰醋酸中肾上腺素的现场光谱电化学行为   总被引:1,自引:0,他引:1  
采用以石墨为工作电极的长光程薄层光谱电化学池, 用循环伏安法、恒电位光谱法和单电位跃计时吸光度法,研究了肾上腺素(AD)在冰醋酸介质中的电子转移反应。结果表明:AD在冰醋酸介质中电氧化反应产物为肾上腺素醌,还原态和氧化态的扩散系数分别为 3. 98×10-6 cm2 /s和 3. 90×10-6 cm2 /s;电极反应的式量电极电位为 0. 632V(vs.SCE),式量异相电子转移速率常数为 7. 30×10-4cm/s;传递系数为 0. 15。  相似文献   

11.
In the reaction of the N-substituted diethanolamines (H(2)L(1-3)) (1-3) with calcium hydride followed by addition of iron(III) or indium(III) chloride, the iron wheels [Fe(6)Cl(6)(L(1))(6)] (4) and [Fe(6)Cl(6)(L(2))(6)] (6) or indium wheels [In(6)Cl(6)(L(1))(6)] (5), [In(6)Cl(6)(L(2))(6)] (8) and [In(6)Cl(6)(L(3))(6)] (9) were formed in excellent yields. Exchange of the chloride ions of 6 by thiocyanate ions afforded [Fe(6)(SCN)(6)(L(2))(6)] (7). Whereas the structures of 4, 5 and 7 were determined unequivocally by single-crystal X-ray analyses, complexes 8 and 9 were characterised by NMR spectroscopy. Contrary to what is normally presumed, the scaffolds of six-membered metallic wheels are not generally rigid, but rather undergo nondissociative topomerisation processes. This was shown by variable temperature (VT) (1)H NMR spectroscopy for the indium wheel [In(6)Cl(6)(L(1))(6)] (5) and is highlighted for the enantiotopomerisation of one indium centre [ 1/6[S(6)-5]<==>[1/6[S(6)-5']]. The self-assembly of metallic wheels, starting from diethanolamine dendrons, is an efficient strategy for the convergent synthesis of metallodendrimers.  相似文献   

12.
The halogenated benzenes C(6)HF(5), 2,4,6-C(6)H(3)F(3), 2,3,5,6-C(6)H(2)F(4), C(6)F(6), C(6)Cl(6), C(6)Br(6), and C(6)I(6) were converted into their corresponding cation radicals by using various strong oxidants. The cation-radical salts were isolated and characterized by electron paramagnetic resonance (EPR) spectroscopy and by single-crystal X-ray diffraction. The thermal stability of the cation radicals increased with decreasing hydrogen content. As expected, the cation radicals [C(6)HF(5)](+) and 2,3,5,6-[C(6)H(2)F(4)](+) had structures with the same geometry as C(6)HF(5) and 2,3,5,6-[C(6)H(2)F(4)]. In contrast, the cation radicals [C(6)F(6)](+), [C(6)Cl(6)](+), and possibly also [C(6)Br(6)](+) exhibited Jahn-Teller-distorted geometries in the crystalline state. In the case of C(6)F(6)(+)Sb(2)F(11)(-), two low-symmetry geometries were observed in the same crystal. Interestingly, the structures of the cation radicals 2,4,6-[C(6)H(3)F(3)](+) and C(6)I(6)(+) did not exhibit Jahn-Teller distortions. DFT calculations showed that the explanation for the lack of distortion of these cations from the D(3h) or D(6h) symmetry of the neutral benzene precursor was different for 2,4,6-[C(6)H(3)F(3)](+) than for [C(6)I(6)](+).  相似文献   

13.
The reaction of [(eta(6)-arene)RuCl(2)](2) (arene = C(6)Me(6), 1,4-MeC(6)H(4)CHMe(2)) with a large excess of the dianion of bis(2-mercaptoethyl) sulfide, (HSCH(2)CH(2))(2)S, obtained from deprotonation of the dithiol with freshly prepared NaOMe, gives the deep red, monomeric complexes [(eta(6)-arene)Ru(eta(3)-C(4)H(8)S(3))] (arene = C(6)Me(6) (5), 1,4-MeC(6)H(4)CHMe(2) (6)) in which the dianion is bound to the metal atom through one thioether and two thiolate sulfur atoms. Complex 5 reacts with [(eta(6)-C(6)Me(6))RuCl(2)](2) (4) in a 2:1 mole ratio to give a quantitative yield of the chloride salt of a binuclear cation [((eta(6)-C(6)Me(6))Ru)(2)Cl(mu(2)-eta(2):eta(3)-C(4)H(8)S(3))](+) (7) in which the thiolate sulfur atoms of the [(eta(6)-C(6)Me(6))Ru(eta(3)-C(4)H(8)S(3))] group bridge to a (eta(6)-C(6)Me(6))RuCl unit. This compound is also obtained directly from the reaction of 4 with the dithiolate, if the Ru dimer is used in large excess. The binuclear complex [((eta(6)-C(6)Me(6))Ru)(2)(MeCN)(mu(2)-eta(2):eta(3)-C(4)H(8)S(3))](PF(6))(2).MeCN, (9)(PF(6))(2).MeCN, is obtained by treatment of (7)Cl with NH(4)PF(6) in acetonitrile. Protonation of 5 with HCl gave the mono- and diprotonated derivatives viz. [(eta(6)-C(6)Me(6))Ru(eta(3)-C(4)H(9)S(3))]Cl, (8)Cl, and [(eta(6)-C(6)Me(6))Ru(eta(3)-C(4)H(10)S(3))]Cl(2), (10)Cl(2), respectively. The reaction of 5 with methyl iodide gives both the mono- and di-S-methylated derivatives. Treatment of 5 with dibromoalkanes, Br(CH(2))(n)Br (n = 1-5), effects ring closure to give the (eta(6)-C(6)Me(6))Ru dications containing the trithia mesocyclic zS3 (z = 8-12) ligands, isolated as their PF(6) salts. The X-ray crystal structures of 5, 6, the solvates of (7)Cl and (9)(PF(6))(2), and the trithia mesocyclic Ru complexes (eta(6)-C(6)Me(6))Ru(zS3)(PF(6))(2) (z = 8-11) are reported.  相似文献   

14.
We analyzed chemical bonding in low-lying isomers of the recently computationally predicted B(6)H(6)Li(6) molecule. According to our calculations the benzene-like B(6)H(6)Li(6) (D(2h), (1)A(1g)) arachno structure with the planar aromatic B(6)H(6)(6-) anion is the most stable one. A nido isomer with two aromatic B(6)H(6)(4-) (pentagonal pyramid) and Li(3)(+) (triangular) moieties, which can be considered as derived from the global minimum structure through a two-electron intramolecular transfer from B(6)H(6)(6-) to three Li(+) cations, was found to be 10.7 kcal/mol higher in energy. A closo isomer with three aromatic moieties (octahedral B(6)H(6)(2-) and two Li(3)(+)) was found to be 31.3 kcal/mol higher in energy than the global minimum. Another isomer with three aromatic moieties (two B(3)H(3)(2-) and Li(3)(+)) was found to be substantially higher in energy (74.4 kcal/mol). Thus, the intramolecular electron transfers from the highly charged B(6)H(6)(6-) anion to cations are not favorable for the B(6)H(6)Li(6) molecule, even when a formation of three-dimensional aromatic B(6)H(6)(2-) anion and two sigma-aromatic Li(3)(+) cations occurs in the closo isomer.  相似文献   

15.
M(eta(6)-arene)(2) species (M = Cr, arene = 1,3,5-Me(3)C(6)H(3); M = Mo, arene = 1,3,5-Me(3)C(6)H(3), 1,3,5-(i)Pr(3)C(6)H(3)), have been prepared by a modified Fischer-Hafner synthesis or by metal vapour techniques. The reaction of Cr(eta(6)-1,3,5-Me(3)C(6)H(3))(2) with the fulvene derivatives pentacarbomethoxycyclopentadiene (pcmcpH), 1-benzoyl-6-hydroxy-6-phenylfulvene (dbcpH), or 1-benzoyl-3-nitro-6-hydroxy-6-phenylfulvene (dbncpH) proceeds with evolution of dihydrogen and formation of the ionic derivatives [Cr(eta(6)-1,3,5-Me(3)C(6)H(3))(2)][E], where E = pcmcp, dbcp, or dbncp. Mo(eta(6)-arene)(2) derivatives (arene = toluene, 1,3,5-Me(3)C(6)H(3), 1,3,5-(i)Pr(3)C(6)H(3)) are oxidized to [Mo(eta(6)-arene)(2)](+) by pcmcpH. The crystal and molecular structures of [M(eta(6)-1,3,5-R(3)C(6)H(3))(2)][pcmcp] (M = Cr, R = Me; M = Mo, R = Me, (i)Pr) have been solved by X-ray single crystal diffraction.  相似文献   

16.
Reactions of C(6)H(5)Li and 4-CH(3)C(6)H(4)Li with halides of Ti, Ir, Hf, and Nb lead to the formation of homoleptic organometallic anions of these metals. Owing to their thermal instability and their sensitivity towards H(2) O and O(2) , these compounds are characterized by single-crystal structure determinations at low temperature, whereas other physical data could only be obtained occasionally. Three pentacoordinate complex anions [Ti(C(6)H(5))(5)](-), [Ti(4-CH(3)C(6)H(4))(5)](-), and [Zr(C(6)H(5))(5)](-) have square-pyramidal structures that display only slight deviations from the ideal geometry, in contrast to the already known structures of [Ti(CH(5))(5)](-). The hexacoordinate complex anions [Zr(C(6)H(5))(6)](2-), [Zr(4-CH(3)C(6)H(4))(6)](2-), [Nb(C(6)H(5))(6)](2-), and [Nb(4-CH(3)C(6)H(4))(6)](2-) all have trigonal-prismatic structures, in accord with the known hexamethyl complex dianions. In contrast, the hexacoordinate complex anion [Hf(C(6)H(5))(6)](2)(-) has an octahedral or close to octahedral structure, in contrast to the known trigonal-prismatic structures of [Ta(C(6)H(5))(6)](-) and [Ta(4-CH(3)C(6)H(4))(6) (-). A qualitative explanation for this structural variability is given.  相似文献   

17.
The products obtained in the reaction of cobalt atoms in neat benzene or in a benzene/argon mixture at low temperature have been reinvestigated. At least three cobalt-containing species were detected by IR, namely, Co(C(6)H(6)), Co(C(6)H(6))(2), and Co(x)(C(6)H(6)), x>1. The IR bands were assigned to these complexes by monitoring their behavior as a function of (a) Co and C(6)H(6) concentration, (b) isotopic substitution, and (c) photoirradiation. We were able to analyze the sample in neat benzene by both electron paramagnetic resonance (EPR) and IR spectroscopy and to determine the magnetic parameters (g tensor and Co hyperfine interaction) for the Co(C(6)H(6))(2) sandwich compound. The large number of fundamental bands observed in the IR spectrum of Co(C(6)H(6))(2), the absorption pattern observed in the Co-ring stretching region of the IR spectrum of the mixed complex, Co(C(6)H(6))(C(6)D(6)) and the orthorhombic g-values extracted from the EPR spectrum are most consistent with nonequivalent benzene ligands in Co(C(6)H(6))(2), i.e., C(s) symmetry. A bonding scheme consistent with both the EPR and IR data for Co(C(6)H(6))(2) is discussed.  相似文献   

18.
Two new anions [Nb(6)F(i)(6)X(i)(6)(NCS)(a)(6)](4-)(X = Br, I) based on octahedral niobium clusters with edge-bridging F ligands have been prepared by reaction of Cs(3)Nb(6)F(6)Br(12) and Cs(4)Nb(6)F(8.5)I(9.5) with aqueous solution of KSCN. The anions were isolated as (Et(4)N)(6)[Nb(6)F(6)Br(6)(NCS)(6)]Br(2) (1)and Cs(1.6)K(2.4)[Nb(6)F(6)I(6)(NCS)(6)] (2) salts.  相似文献   

19.
Solid-state decomposition of [V3O(O2CPh)6(H2O)3]Cl at 300 degrees C followed by alcoholysis of the product gives the new vanadium complexes [V6O6(PhCO2)6(CH3O)6(CH3OH)3] (1), [V6O6(PhCO2)6(C2H5O)6(C2H5OH)3] (2), [V6O6(PhCO2)6(C3H7O)6(C3H7OH)3] (3), [V6O6(PhCO2)6(C4H9O)6(C4H9OH)3] (4) and [V4O4(OCH3)6(O2CPh)2(HOCH3)2] (5). Complexes 2, 3 and 5 have been crystallographically characterised. DC magnetic susceptibility studies on complex shows antiferromagnetic coupling leading to a S = 0 spin ground state.  相似文献   

20.
Diuranium μ-η(6),η(6)-arene complexes supported by ketimide ligands were synthesized and characterized. Disodium or dipotassium salts of the formula M(2)(μ-η(6),η(6)-arene)[U(NC(t)BuMes)(3)](2) (M = Na or K, Mes = 2,4,6-C(6)H(2)Me(3)) and monopotassium salts of the formula K(μ-η(6),η(6)-arene)[U(NC(t)BuMes)(3)](2) (arene = naphthalene, biphenyl, trans-stilbene, or p-terphenyl) were both observed. Two different salts of the monoanionic, toluene-bridged complexes are also described. Density functional theory calculations have been employed to illuminate the electronic structure of the μ-η(6),η(6)-arene diuranium complexes and to facilitate the comparison with related transition-metal systems, in particular (μ-η(6),η(6)-C(6)H(6))[VCp](2). It was found that the μ-η(6),η(6)-arene diuranium complexes were isolobal with (μ-η(6),η(6)-C(6)H(6))[VCp](2) and that the principal arene-binding interaction was a pair of δ bonds (total of 4e) involving both metals and the arene lowest unoccupied molecular orbital. Reactivity studies have been carried out with the mono- and dianionic μ-η(6),η(6)-arene diuranium complexes, revealing contrasting modes of redox chemistry as a function of the system's state of charge.  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