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 109 毫秒
1.
设计用4种自由基自旋中心连接在耦合单元喹啉的不同位置上的双自由基体系,用AM1-CI方法计算的结果表明:双自由基连接的位置不同对体系耦合作用的影响符合双自由基之间磁性耦合的拓扑规则,即共轭体系中,两个自由基之间以偶数个C(或N)原子耦合,体系具有低自旋基态,表现反铁磁耦合;两个自由基之间以奇数个C(或N)原子耦合,体系具有高自旋基态,表现铁磁耦合.当双自由基连接在喹啉的相邻奇数个C或N原子位置时,体系具有高自旋基态,表现铁磁耦合.  相似文献   

2.
采用量子化学abinitio法对具有甲氧基的碳、氧双自由基邻、间、对二甲氧基亚甲基苯及衍生物体系基态自旋耦合规律进行研究,得到非平面共轭体系中自由基之间磁性耦合的拓朴规则:共轭体系中,两个自由基之间以偶数个碳原子耦合,则有效交换积分J~i~j<0,体系具有低自旋基态;两个自由基之间以奇数个碳原子耦合,则J~i~j>0,体系具有高自旋基态。自由基性质对自旋耦合的影响较大,正离子自由基间磁性耦合能力较强,这些结论为有机磁性材料的分子设计与实验合成提供了理论依据。  相似文献   

3.
采用量子化学abinitio法和密度泛函方法对不同取代位置的嘧啶自旋耦合规律进行研究 .两种方法比较 ,用UHF方法计算导致自旋污染严重 ,而用UB3LYP方法计算 ,自旋污染则减少了许多 .计算结果得到了双自由基之间磁性耦合的拓扑规则 :共轭体系中 ,两个自由基之间以偶数个碳 (或氮 )原子耦合 ,则有效交换积分Jij<0 ,体系具有低自旋基态 ,表现为反铁磁耦合 ;两个自由基之间以奇数个碳 (或氮 )原子耦合 ,则有效交换积分Jij>0 ,体系具有高自旋基态 ,表现为铁磁耦合 .自由基性质和铁磁耦合单元的不同位置对自旋耦合的影响较大 ,这些结论为有机磁性材料的分子设计与实验合成提供了理论依据 .  相似文献   

4.
仇永清  赵成大 《化学学报》1996,54(12):1177-1180
采用量子化学从头算UHF方法对系列有机双自由基体系的基态自旋耦合规律进行研究, 进一步证实了自由基之间在共轭体系中出现铁磁性耦合的拓扑规则,统一了关于自由基耦合规律的几种解释, 为有机磁性材料的实验合成提供了理论指导。  相似文献   

5.
采用量子化学从头算UHF方法对系列有机双自由基体系的基态自旋耦合规律进行研究, 进一步证实了自由基之间在共轭体系中出现铁磁性耦合的拓扑规则,统一了关于自由基耦合规律的几种解释, 为有机磁性材料的实验合成提供了理论指导。  相似文献   

6.
仇永清  赵成大 《化学学报》1997,55(2):129-132
采用量子化学从头算UHF方法, 对平面型双卡宾及双氮宾体系的基态自旋情况进行研究。结合前面的分析结果, 进一步探讨了多自由基体系基态自旋的耦合规律, 为有机磁性体的分子设计提供了可靠的理论依据。  相似文献   

7.
赵红梅  傅强  仇永清  陈丽莉 《化学学报》2001,59(11):1853-1857
采用量子化学abinitio方法,讨论了2,2'-联吡啶同自旋双自由基体系构象变化对铁磁耦合的影响。结果表明,在各种构象下,体系的磁性耦合符合自旋极化规则;对于·CH2,·MH2^+两种自由基磁性耦合性质是机同的,只影响到体系磁性耦合的强度,这一结论为有机磁性材料的分子设计提供了有益的信息。  相似文献   

8.
采用量子化学ab initio方法,讨论了2,2′-联吡啶同自旋双自由基体系构象变化对铁磁耦合的影响.结果表明,在各种构象下,体系的磁性耦合符合自旋极化规则;对于·CH2,·NH2+两种自由基磁性耦合性质是相同的,只影响到体系磁性耦合的强度,这一结论为有机磁性材料的分子设计提供了有益的信息.  相似文献   

9.
孙友敏  刘成卜  王若曦 《化学学报》2004,62(23):2303-2307
用密度泛函理论结合对称性破损态方法对氮氧双自由基以及铜(Ⅱ)-氮氧自由基配合物的磁耦合常数进行了计算.结果表明铜(Ⅱ)-氮氧自由基配合物为铁磁耦合.对配合物磁轨道进行了分析,表明体系的铁磁耦合作用主要来自于Cu离子的轨道与自由基的π*轨道正交.自旋密度分布分析显示:在氮氧自由基与金属铜两个自旋耦合片的自旋耦合主要来自于中心Cu离子的轨道电子向氮氧自由基上的π*轨道的电子转移,这一电子特征的变化引起的自旋离域在Cu离子和氮氧自由基片的铁磁耦合中起到了重要的作用.  相似文献   

10.
利用AM1-CI方法计算了构象对邻、间、对二取代苯双自由基体系基态自旋多重度及其稳定性的影响. 结果表明单 - 三重态能量差(△ES-T)和部分占据分子轨道的能量劈裂(△EPOMO)随自由基与苯环间的二面角而变化. 当二面角接近90°时, 分子具有平面或近平面构象时强的铁磁或反铁磁耦合单元, 由于具有近简并的高自旋和低自旋基态, 而变成弱的反铁磁或铁磁耦合单元. 由此提出为获得具有稳定高自旋基态的高自旋分子, 实验上应尽量避免选用强烈扭曲的邻、对苯分子构象.  相似文献   

11.
Henze and Blair [2] have successfully derived algorithms for enumerating the number of constitutional isomers of aliphatic compounds using alkyl radicals; however, the algorithms cannot be extended to enumerate constitutional isomers of cyclic compounds. Similarly, Read [4] has advocated the use of alkyl biradicals to enumerate constitutional isomers of aliphatic compounds, but not cyclic compounds. Apparently, the use of alkyl biradicals in the enumeration of constitutional isomers of cyclic compounds has been neglected. In this communication, an algorithm using alkyl biradicals to enumerate the number of constitutional isomers of cyclic compounds, namely alkylcyclopropanes, is described. This revised version was published online in July 2006 with corrections to the Cover Date.  相似文献   

12.
Vertical electron affinities (EA) are predicted for the lowest energy singlet states of the 21 didehydroquinolinium cation isomers and the 21 didehydroisoquinolinium cation isomers, as well as the doublet states of the seven dehydroquinolinium cation isomers, the seven dehydroisoquinolinium cation isomers, the seven N-methyldehydroquinolinium cations, and the seven N-methyldehydroisoquinolinium cations, by using density functional theory. For the monoradicals, the calculated EA of the radical site depends only on the distance from the (formally charged) nitrogen atom, and is reduced by 0.14-0.24 eV when the NH+ group is replaced with an NCH3+ group. Nearly all of the calculated EAs for the ortho biradicals are lower (by 0.04-0.72 eV) than those for either of the corresponding monoradicals. For the meta biradicals, the calculated EAs lie either between the EAs of the corresponding monoradicals or higher (by 0.07-0.58 eV), and they are extremely sensitive to the separation (distance) between the two dehydrocarbon atoms. For the biradicals that do not have either an ortho or meta relationship the calculated EAs are all higher (by 0.02-1.93 eV) than those for either of the corresponding monoradicals. The EAs are examined to gain insight into the nature of inductive/field and resonance effects that influence the electrophilicity of the radical site(s), which is a major factor controlling the reactivity of these types of (bi)radicals.  相似文献   

13.
CASPT2//CASSCF and B3LYP methodologies have been used to study the excited-state properties and photochemical isomerizations of p-, m-, and o-methylbenzonitriles and methylanisoles. Calculations show that the biradical mechanism is the most favored channel for the photoinduced interconversion of p-, m- and o-methylbenzonitriles, both dynamically and thermodynamically. The formation of biradical as a key intermediate is highly selective, and only the biradicals with a turned-up cyano-substituted carbon are involved in photoisomerization. Methylanisole isomers are inactive relative to methylbenzonitriles at 254 nm. Such remarkable activity difference between methylbenzonitrile and methylanisole in photochemistry arises from the accessibility of the S1/S0 conical intersection as well as the stability of prefulvene biradicals. For methylanisoles, the S1/S0 precursor and the reactive biradicals are inaccessible at 254 nm, which should be the origin of inactivity. The results suggest that the conical intersection accessibility plays a crucial role in the photochemistry of substituted benzenes at 254 nm.  相似文献   

14.
The magnetic coupling in organic biradicals has been analyzed by means of ab initio wave function-based methods. Attention is focused on the coupling between the spin moments localized on the NO-groups in meta and para phenylene-bridged nitroxides, and bis(nitronyl) nitroxide and bis(imino) nitroxide biradicals. The leading mechanisms governing the coupling have been isolated by means of class-partitioned CI calculations. It was found that the mechanisms of the coupling in the para and meta phenylene-bridged nitroxides are similar to that found in transition metal complexes, while for the other biradicals the dominance of other mechanisms (like the spin polarization) imposes restrictions on the computational strategy to be followed to best estimate the coupling.  相似文献   

15.
Four pi,pi-biradicals, 2,6-dimethylenepyridinium and the novel isomers N-(3-methylenephenyl)-3-methylenepyridinium, N-phenyl-3,5-dimethylenepyridinium, and N-(3,5-dimethylenephenyl)pyridinium ions, were generated and structurally characterized in a Fourier transform ion cyclotron resonance mass spectrometer. Their gas-phase reactivity toward various reagents was compared to that of the corresponding monoradicals, 2-methylenepyridinium, N-phenyl-3-methylenepyridinium, and N-(3-methylenephenyl)pyridinium ions. The biradicals reactivity was found to reflect their predicted multiplicity. The 2,6-dimethylenepyridinium ion, the only biradical in this study predicted to have a closed-shell singlet ground state, reacts significantly faster than the other biradicals, which are predicted to have triplet ground states. In fact, this biradical reacts at a higher rate than the analogous monoradical, which suggests that to avoid the costly uncoupling of its unpaired electrons, the biradical favors ionic mechanisms over barriered radical pathways. In contrast, the second-order reaction rate constants of the isomeric biradicals with triplet ground states are well approximated by those of the analogous monoradicals, although the final reaction products are sometimes different. This difference arises from rapid radical-radical recombination of the initial monoradical reaction products. The overall reactivity toward the hydrogen-atom donors benzeneselenol and tributylgermanium hydride is significantly greater for the radicals with the charged site in the same ring system as the radical site. This finding indicates that polar effects play an important role in controlling the reactivity of pi,pi-biradicals, just as has been demonstrated for sigma,sigma-biradicals.  相似文献   

16.
A simple and fast algorithm for enumerating alkene isomers is described. This algorithm is based on a 2-step hypothetical chemical reaction. In the first step, an alkane molecule loses two protons becoming an alkyl biradical. In the second step, two alkyl biradicals react to form an alkene molecule. By using two sequential recursive algorithms, the number of computational terms in the enumeration process is much less than when using the Henze–Blair algorithm. By using this simple algorithm, the numbers of constitutional isomers of alkenes and alkynes for carbon content greater than 30 are easily enumerated. In addition, through this algorithm, a mathematical relationship between the number of constitutional isomers of alkenes and alkynes has been established for the first time. This revised version was published online in July 2006 with corrections to the Cover Date.  相似文献   

17.
Four biradicals of differing chain lengths have been prepared by spin-labelling the sulphhydryl groups of dithiothreitol. Temperature-dependent EPR spectra of these biradicals indicate that the dynamics of biradicals is associated with both fast and slow exchanges of conformations. The three-conformational model incorporating cage configuration is found to be more appropriate. Certain thermodynamic parameters of conformationalchanges have been calculated from the experimental EPR parameters, obtained by the use of simulation procedures developed recently.  相似文献   

18.
Unexpectedly, the 5-dehydroquinoline radical cation was formed in the gas phase from the 5-iodo-8-nitroquinolinium cation upon ion-trap collision-activated dissociation. This reaction involves the cleavage of a nitro group to generate an intermediate monoradical, namely, the 8-dehydro-5-iodoquinolinium cation, followed by rearrangement through abstraction of a hydrogen atom from the protonated nitrogen atom by the radical site. Dissociation of the rearranged radical cation through elimination of an iodine atom generates the 5-dehydroquinoline radical cation. The mechanism was probed by studying isomeric biradicals and performing quantum chemical calculations. The 5-dehydroquinoline radical cation showed greater gas-phase reactivity toward dimethyl disulfide, cyclohexane, and allyl iodide than the isomeric 5,8-didehydroquinolinium cation, which is more reactive than the isomeric 5,8-didehydroisoquinolinium cation studied previously. All three isomers have a 1,4-biradical topology. The order of reactivity is rationalized by the vertical electron affinities of the radical sites of these biradicals instead of their widely differing singlet–triplet splittings.  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