首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到18条相似文献,搜索用时 31 毫秒
1.
1,4-与1,2-二氢NADH模型物反应活性的比较   总被引:1,自引:1,他引:0  
合成了两类NADH模型物: p-G-1,4-2H-PNAH和p-G-1,2-2H-PNAH, 测得了它们在不同温度下的乙腈中分别与四氯苯醌及N,N,N',N'-四甲基对苯二胺自由基正离子(TMPA+·)反应的速率常数及反应的活化参数(ΔH≠, ΔS≠, ΔG≠). 动力学研究结果表明: 1,2-2H-PNAH和1,4-2H-PNAH的相对反应活性可以通过远位取代基调节; 1,2-2H-PNAH和1,4-2H-PNAH与四氯苯醌的反应是决速步骤的熵控反应, 而1,2-2H-PNAH和1,4-2H-PNAH与TMPA+·的反应决速步骤则是由焓和熵共同控制的反应; 通过lnk2~σ的相关分析可以看出1,2-2H-PNAH和1,4-2H-PNAH与四氯苯醌及TMPA+·反应决速步骤中反应中心是正电荷增加的过程, 而且1,2-2H-PNAH的取代基效应大于1,4-2H-PNAH的取代基效应.  相似文献   

2.
刘有成  李劲 《有机化学》2001,21(11):898-903
对本课题组近年来研究的辅酶NADH模型物还原活化烯烃的反应机理进行了综述。对于辅酶模型物还原2-溴-1-苯基亚乙基丙二腈类化合物的反应,依赖辅酶模型物和底物的结构,反应可以按一步的负氢转移机理或按电子转移机理进行。用手性辅酶模型物进行这一反应,可得到具有中等光学活性的环丙烷衍生物。实验结果表明辅酶模型物BNAH与1,1-二苯基-2,2-二硝基乙烯的反应的过渡态具有部分双自由基和部分共价键形成的特征,为Pross-Shaik“曲线交叉模型”所预测的“中间机理”提供了直接的证据。BNAH与9-亚芴基丙二腈的反应经历电子转移和负电荷在9-位碳上的碳负离子中间体,动力学同位素效应为2.6。  相似文献   

3.
辅酶 NADH模型物还原反应机理研究长期以来一直是物理有机化学领域令人感兴趣的课题 [1~ 3 ] ,其中对于反应动力学的研究因其对推论反应机理的重要性而受到重视 .我们 [4 ,5] 曾报道过辅酶 NADH模型物 N -苄基 - 1 ,4-二氢烟酰胺 (BNAH)与 2 -溴 - 1 -苯基亚乙基丙二腈 (BPM)的还原脱溴环丙烷化反应 ,通过紫外 -可见光谱法测定了该反应的反应速率常数、同位素效应等动力学方面的结果 .然而紫外 -可见光谱法本身有一定的缺陷 ,例如 ,由于反应物在紫外 -可见光谱中的吸收峰可能互相重叠 ,有时会观察不到产物和中间体的形成 ;或者有时…  相似文献   

4.
 采用原位制备的铱双膦(或膦氮)配合物在碘存在下催化2-取代-1,2-二氢喹啉、2-甲基-2,3-二氢吲哚、1,4-二氢吡啶及3,4-二氢异喹啉等化合物的脱氢芳构化反应, 并考察了不同金属前体、配体、催化剂用量、溶剂和碘等因素对反应速率和选择性的影响. 结果表明,原位制备的[Ir(COD)Cl]2/(±)-MeO-Biphep在碘的存在下催化2-取代-1,2-二氢喹啉的脱氢芳构化反应速率快, 选择性好,催化剂的用量少; 对1,4-二氢吡啶和2,3-二氢吲哚的催化脱氢芳构化反应则须在高温下进行; 而对 3,4-二氢异喹啉, 即使在加热回流条件下也只有不到5%的转化率. 催化体系中碘的存在可以明显提高反应速率.  相似文献   

5.
4位取代的Hantzsch酯(HEH)衍生物在2-硝基-2-亚硝基丙烷的氧化下生成相应的吡啶类化合物. 将N-氘代1,4-二氢Hantzsch酯(N-d-HEH)和4,4'-双氘代1,4-二氢Hantzsch酯(4,4'-2d-HEH)分别代替HEH与2-硝基-2-亚硝基丙烷反应, 得到的同位素效应常数分别为1.03(kN-H/kN-D)和1.78(kC4-H/kC4-D), 表明1,4-二氢Hantzsch酯中4位上氢原子所涉及的C4-H键的断裂发生在反应的决速步骤中或在决速步骤之前, 而1位上氢原子所涉及键的断裂则不在决速步骤中. 由4位取代的HEH酯衍生物的氧化电位与2-硝基-2-亚硝基丙烷的还原电位可在热力学上判断该反应不是由电子转移引发的. 向反应体系中加入单电子转移抑制剂对二硝基苯, 反应未受到明显抑制, 进一步证明了上述推断. 据此推测, 反应可能是通过NO+直接对HEH酯上氮原子的亲电历程引发的.  相似文献   

6.
1,6-二氢-S-四嗪与异氰酸酯催化重排反应的研究   总被引:5,自引:0,他引:5  
3,6-二甲基-1,6-二氢-S-四嗪与异氰酸酯反应,在N′,N-二甲基苯胺催化下生成N-苯基-3,6-二甲基-1,6-二氢-S四嗪-1-甲酰胺,在4-二甲胺基吡啶(DMAP)催化则生成一类新化合物.与14种不同的取代苯基异氰酸酯反应,生成14种新化合物,产率中等.化合物3i经X射线单晶衍射分析表明其结构为N,N′-双(邻甲苯基)-3,6-二甲基-1,4-二氢-S-四嗪-1,4-二甲酰胺,这表明发生了重排.据1H NMR测定结果,对反应机理作了初步讨论.14种化合物体外抗癌活性检测表明部分化合物有强的生物活性.  相似文献   

7.
芳基二氢萘类衍生物是许多生物活性的天然产物以及药物的常见结构单元,其合成一直都受到化学家们的关注.传统的1,2-二氢-1-芳基萘骨架化合物的构筑大都需要进行底物的预官能团化,在高温条件下进行,且产物的选择性较差,因此发展一种简单温和的制备方法很有必要.最近兴起的可见光催化因具有条件温和、环境友好等特点而成为了合成化学家的研究热点.近期研究发现,在可见光作用下利用吖啶光敏剂的强氧化能力,可以实现苯乙烯的加成.但此类反应需要当量的氧化剂或氢原子转移试剂,容易导致苯乙烯的二聚环合产物的进一步氧化或还原.我们在前期发展的"放氢交叉偶联"反应的基础上,利用吖啶光催化和钴肟催化的协同作用,实现了苯乙烯的放氢二聚反应,在室温下高效构筑了1,2-二氢-1-芳基萘骨架,反应条件温和,底物脱除的电子和质子在钴肟催化剂作用下以氢气的形式释放,反应具有中等及以上的收率.本文以苯乙烯为模型底物,吖啶为光敏剂,钴肟配合物为质子还原催化剂,在乙腈溶剂中,蓝色LED灯下光照24 h可以获得56%的产率,对于其它的光敏剂如fac-Ⅰr(ppy)3等则不能催化该反应.通过催化剂种类及用量筛选表明,7 mol%的Co(dmgH_2)pyCl配合物具有最好的反应效果,可以获得72%的收率.控制实验表明,光敏剂、钴肟催化剂和光照都是必须的.通过底物拓展我们发现,烷基、卤素等不同取代基的苯乙烯类化合物均可以获得较好的收率,不同苯乙烯之间也可以发生交叉反应.随后,我们进一步通过光谱和中间体捕获实验对反应机理进行了研究.自由基捕获实验说明反应过程可能涉及自由基历程;光谱淬灭实验表明苯乙烯和Co(dmgH_2)pyCl均可淬灭吖啶的发光,但苯乙烯淬灭吖啶的程度远大于Co(dmgH_2)pyCl淬灭吖啶的程度.在反应时苯乙烯的浓度远大于催化剂的溶度,因此,我们认为激发态吖啶首先与苯乙烯发生反应;可见光照射反应体系1 min后在440–500和550–650 nm处观察到明显的Co~Ⅱ和Co~Ⅰ的吸收峰.基于以上实验结果,我们提出了可能的催化循环:吖啶受光激发到达激发态后,首先与底物苯乙烯发生单电子转移生成苯乙烯正离子自由基和吖啶阴离子自由基Acr~·-Mes,Acr~·-Mes还原Co(dmgH_2)pyCl生成Co ~Ⅱ中间体,从而回到基态完成光催化循环.苯乙烯正离子自由基与另一分子苯乙烯加成环合,进而通过芳构化生成自由基中间体,再与Co Ⅱ作用生成目标产物1,2-二氢-1-芳基萘和Co~Ⅰ,Co~Ⅰ通过结合体系中的质子进而释放出氢气回到Co~ Ⅲ从而完成钴肟催化循环.  相似文献   

8.
设计、合成了一类新型的高分子还原试剂--聚苯乙烯固载烟酰胺辅酶模型化合物1-苄基-1,4-二氢烟酰胺(BNAH)4.该还原剂可以在温和条件下有效的还原活化烯烃,而且可以循环再生,但循环再生后有效BNAH含量下降.  相似文献   

9.
辅酶NAD(P)H在生物氧化还原反应中起着重要作用[1].1-苄基-1,4-二氢尼古丁酰胺(BNAH)作为其模型物,被广泛用于物理有机和生物化学的研究之中[2].虽然绝大多数的研究都集中于还原反应机理方面[3,4],BNAH作为还原剂在有机合成中的应用也是值得注意的.我们曾用BNAH还原2-溴-1-苯亚乙基丙二腈及其类似物合成取代环丙烷[5~7],方法简便.五元环结构广泛存在于萜类和甾体等天然产物中.对于茚等苯并五元环结构的合成已有许多方法[8~11]. 其中,2,2-双取代1,2-二氢茚(1)(吸电子取代基)是用邻-二溴甲基苯与丙二腈等活泼亚甲基化合物在DMSO中,NaH存在下双分子缩合制备的[12].  相似文献   

10.
设计并合成了2-[1-(4-(N-苯基-N-丙酰基)氨基)哌啶基]甲基-六氢-1H-1,4-二氮 7.由3经还原、氯取代、胺取代、胺脱苄反应得到了7.7经单酰化反应得到了8.8经酰化反应后得到了11个带有(1-芳乙酰胺基-2-叔氨基)乙烷结构的六氢-1H-1,4-二氮类目标化合物(9a~9g,10a~10d).  相似文献   

11.
Five 1-(p-substituted phenyl)-1,4-dihydronicotinamides (GPNAH-1,4-H(2)) and five 1-(p-substituted phenyl)-1,2-dihydronicotinamides (GPNAH-1,2-H(2)) were synthesized, which were used to mimic NAD(P)H coenzyme and its 1,2-dihydroisomer reductions, respectively. When the 1,4-dihydropyridine (GPNAH-1,4-H(2)) and the 1,2-dihydroisomer (GPNAH-1,2-H(2)) were treated with p-trifluoromethylbenzylidenemalononitrile (S) as a hydride acceptor, both reactions gave the same products: pyridinium derivative (GPNA(+)) and carbanion SH(-) by a hydride one-step transfer. Thermodynamic analysis on the two reactions shows that the hydride transfer from the 1,2-dihydropyridine is much more favorable than the hydride transfer from the corresponding 1,4-dihydroisomer, but the kinetic examination displays that the former reaction is remarkably slower than the latter reaction, which is mainly due to much more negative activation entropy for the former reaction. When the formed pyridinium derivative (GPNA(+)) was treated with SH(-), the major reduced product was the corresponding 1,4-dihydropyridine along with a trace of the 1,2-dihydroisomer. Thermodynamic and kinetic analyses on the hydride transfer from SH(-) to GPNA(+) all suggest that the 4-position on the pyridinium ring in GPNA(+) is much easier to accept the hydride than the 2-position, which indicates that when the 1,4-dihydropyridine is used the hydride donor to react with S, the formed pyridinium derivative GPNA(+) may return to the 1,4-dihydropyridine by a hydride transfer cycle; but when the 1,2-dihydropyridine is used as the hydride donor, the formed pyridinium derivative can not return to the 1,2-dihydropyridine by the hydride reverse transfer from SH(-) to GPNA(+). These results clearly show that the hydride-transfer cycle is favorable for the 1,4-dihydronicotinamides, but unfavorable for the corresponding 1,2-dihydroisomers.  相似文献   

12.
Heterolytic and homolytic bond dissociation energies of the C4-H bonds in ten NADH models (seven 1,4-dihydronicotinamide derivatives, two Hantzsch 1,4-dihydropyridine derivatives, and 9,10-dihydroacridine) and their radical cations in acetonitrile were evaluated by titration calorimetry and electrochemistry, according to the four thermodynamic cycles constructed from the reactions of the NADH models with N,N,N',N'-tetramethyl-p-phenylenediamine radical cation perchlorate in acetonitrile (note: C9-H bond rather than C4-H bond for 9,10-dihydroacridine; however, unless specified, the C9-H bond will be described as a C4-H bond for convenience). The results show that the energetic scales of the heterolytic and homolytic bond dissociation energies of the C4-H bonds cover ranges of 64.2-81.1 and 67.9-73.7 kcal mol(-1) for the neutral NADH models, respectively, and the energetic scales of the heterolytic and homolytic bond dissociation energies of the (C4-H)(.+) bonds cover ranges of 4.1-9.7 and 31.4-43.5 kcal mol(-1) for the radical cations of the NADH models, respectively. Detailed comparison of the two sets of C4-H bond dissociation energies in 1-benzyl-1,4-dihydronicotinamide (BNAH), Hantzsch 1,4-dihydropyridine (HEH), and 9,10-dihydroacridine (AcrH(2)) (as the three most typical NADH models) shows that for BNAH and AcrH(2), the heterolytic C4-H bond dissociation energies are smaller (by 3.62 kcal mol(-1)) and larger (by 7.4 kcal mol(-1)), respectively, than the corresponding homolytic C4-H bond dissociation energy. However, for HEH, the heterolytic C4-H bond dissociation energy (69.3 kcal mol(-1)) is very close to the corresponding homolytic C4-H bond dissociation energy (69.4 kcal mol(-1)). These results suggests that the hydride is released more easily than the corresponding hydrogen atom from BNAH and vice versa for AcrH(2), and that there are two almost equal possibilities for the hydride and the hydrogen atom transfers from HEH. Examination of the two sets of the (C4-H)(.+) bond dissociation energies shows that the homolytic (C4-H)(.+) bond dissociation energies are much larger than the corresponding heterolytic (C4-H)(.+) bond dissociation energies for the ten NADH models by 23.3-34.4 kcal mol(-1); this suggests that if the hydride transfer from the NADH models is initiated by a one-electron transfer, the proton transfer should be more likely to take place than the corresponding hydrogen atom transfer in the second step. In addition, some elusive structural information about the reaction intermediates of the NADH models was obtained by using Hammett-type linear free-energy analysis.  相似文献   

13.
Stefan C. Ritter 《Tetrahedron》2005,61(22):5241-5251
A series of chiral peptides has been prepared, bearing a 1,4-dihydronicotine amide and a zinc cyclen moiety. The metal complex reversibly binds flavins in aqueous solution, while the dihydronicotine amide serves as a NADH model transferring a hydride to the flavin within the assembly. The reaction rate of the redox reaction was monitored and determined by UV spectroscopy. The reaction rates of the substituted compounds were slower if compared to the non-substituted parent compound 1-H, but still show a 30-100 fold rate enhancement compared to the compound missing a flavin binding site. It was anticipated to probe the cryptic stereoselectivity of the hydride transfer from dihydropyridine to flavin. Spectroscopic data indicate that the introduction of deuterium labels upon reduction of the pyridinium salts to 1,4-dihydropyridine in D2O proceeds diastereoselectively, but identical isotope effects on the rate of flavin reduction as with a non-chiral NADH model revealed that the hydride transfer within the assembly proceeds not stereoselective. A more rigid chiral NADH model compound must be prepared to achieve this goal.  相似文献   

14.
Hydride exchange occurs between 3,5 - di(alkoxycarbonyl) - 1,4 - dihydropyridines and their corresponding pyridinium salts. For the case of 1,2,6 - trimethyl - 3,5 - di(ethoxycarbonyl) - 1,4 - dihydropyridine in the presence of the structurally corresponding pyridinium perchlorate, hydride is transferred to the 4-position of the pyridinium salt in a reversible “blind” reaction as revealed by deuterium labeling experiments and to the 2,6-positions irreversibly to afford 1,2,6 - trimethyl - 3,5 - di(ethoxycarbonyl) - 1,2 - dihydropyridine as final product. Removal of the methyl groups at the 2,6-positions, i.e. 1 - methyl - 3,5 - di(methoxycarbonyl) - 1,4 -dihydropyridine and its structurally corresponding pyridium perchlorate, causes hydride transfer to become completely reversible. Substitution of the 4-position with Me, i.e. 1,2,4,6 - tetramethyl - 3,5 - di(methoxycarbonyl) -1,4- dihydropyridine and its corresponding pyridinium perchlorate leads to cessation of hydride transfer: the same is true for the analogous 4-phenyl (and substituted phenyl) compounds. However, these 1,4-dihydropyridines are capable of transferring hydride at reasonable temperatures to less highly substituted pyridinium salts. Activation parameters for some of these hydride transfers have been determined, mechanistic conclusions are presented, and the consequences of these observations for experiments with “model” NADH compounds are discussed.  相似文献   

15.
We consider the controversial conclusions of the charge at the migrating hydrogen in the transition state of hydride‐transfer reactions from CH‐groups to hydride acceptors. Quantum chemical calculations were performed on elementary organic reactions involving carbenium ions, which can be considered as hydride acceptors. We also discuss the biochemical hydride‐transfer reactions in which the coenzyme NADH‐NAD+ plays an important role. With the calculations and the experimental model systems, an answer is given for the stereospecificity of the hydride transfer. Generally, the hydride transfer occurs via a trigonal pyramidal geometry in which the transferred hydride of the CH‐group is located in the axial position. In the case of the coenzyme NADH‐NAD+, the hydride transfer is coupled with an out‐of‐plane orientation of the carboxamide group of the pyridinium moiety, resulting in an increased stereospecificity. © 2004 Wiley Periodicals, Inc. Int J Quantum Chem, 2005  相似文献   

16.
Hydride transfer from dihydronicotinamide adenine dinucleotide (NADH) analogues, such as 10-methyl-9,10-dihydroacridine (AcrH 2) and its derivatives, 1-benzyl-1,4-dihydronicotinamide (BNAH), and their deuterated compounds, to non-heme oxoiron(IV) complexes such as [(L)Fe (IV)(O)] (2+) (L = N4Py, Bn-TPEN, and TMC) occurs to yield the corresponding NAD (+) analogues and non-heme iron(II) complexes in acetonitrile. Hydride transfer from the NADH analogues to p-chloranil (Cl 4Q) also occurs to produce the corresponding NAD (+) analogues and the hydroquinone anion (Cl 4QH (-)). The logarithms of the observed second-order rate constants (log k H) of hydride transfer from NADH analogues to non-heme oxoiron(IV) complexes are linearly correlated with those of hydride transfer from the same series of NADH analogues to Cl 4Q, including similar kinetic deuterium isotope effects. The log k H values of hydride transfer from NADH analogues to non-heme oxoiron(IV) complexes are also linearly correlated with those of deprotonation of the radical cations of NADH analogues. Such linear correlations indicate that overall hydride-transfer reactions of NADH analogues to both non-heme oxoiron(IV) complexes and Cl 4Q occur via electron transfer from NADH analogues to the oxoiron(IV) complexes, followed by rate-limiting deprotonation from the radical cations of NADH analogues and subsequent rapid electron transfer from the deprotonated radicals to the Fe(III) complexes to yield the corresponding NAD (+) analogues and the Fe(II) complexes. The electron-transfer pathway was accelerated by the presence of perchloric acid, and the resulting radical cations of NADH analogues were detected by electron spin resonance spectroscopy and UV-vis spectrophotometry in the acid-promoted hydride-transfer reactions from NADH analogues to non-heme oxoiron(IV) complexes. This result provides the first direct evidence that a hydride transfer from NADH analogues to non-heme oxoiron(IV) complexes proceeds via an electron-transfer pathway.  相似文献   

17.
We report the application of 4‐nitrophenyl diazonium modified electrodes towards the electrochemical detection of NADH. Selective activation of individual electrodes on a 5 element array by electro‐addressable conversion of nitro groups to amines and subsequent EDC/NHS crosslinking to the NADH oxidant, pyrroloquinoline quinone (PQQ), is demonstrated. Inactivated electrodes retained nitro functionality and were protected against non‐specific adsorption and mild chemical reactions. Electrodeposition conditions were used to control nitrophenyl film thickness and showed that while increased film thickness leads to greater functionalization density of PQQ, it also results in decreased electron transfer kinetics. The electrodeposition protocol can therefore serve as a method to control electrode functionalization density and film electron transfer kinetics. We believe this simple technique for selective electrode functionalization may facilitate the development of next generation multianalyte electrochemical sensors.  相似文献   

18.
Reactions of the β-diketiminato n-butyl magnesium complex, [HC{(Me)CN(2,6-(i)Pr(2)C(6)H(3))}(2)Mg(n)Bu], with a range of substituted pyridines and fused-ring quinolines in the presence of PhSiH(3) has been found to result in dearomatisation of the N-heterocyclic compounds. This reaction is proposed to occur through the formation of an unobserved N-heterocycle-coordinated magnesium hydride and subsequent hydride transfer via the C2-position of the heterocycle prior to hydride transfer to the C4-position and formation of thermodynamically-favoured magnesium 1,4-dihydropyridides. This reaction is kinetically suppressed for 2,6-dimethylpyridine while the kinetic product, the 1,2-dihydropyridide derivative, was isolated through reaction with 4-methylpyridine (4-methylpyridine), in which case the formation of the 1,4-dihyropyridide is prevented by the presence of the 4-methyl substituent. X-ray structures of the products of these reactions with 4-methylpyridine, 3,5-dimethylpyridine and iso-quinoline comprise a pseudo-tetrahedral magnesium centre while the regiochemistry of the particular dearomatisation reaction is determined by the substitution pattern of the N-heterocycle under observation. The compounds are all air-sensitive and exposure of the magnesium derivatives of dearomatised pyridine and 4-dimethylaminopyridine (DMAP) to air resulted in ligand rearomatisation and the formation of dimeric μ(2)-η(2)-η(2)-peroxomagnesium compounds which have also been subject to analysis by single crystal X-ray diffraction analysis. An unsuccessful extension of this chemistry to N-heterocycle hydrosilylation is suggested to be a consequence of the low basicity of the silane reagent in comparison to the pyridine substrates which effectively impedes any further interaction with the magnesium centres.  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