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 gas-phase acidities of the six dimethylphenol isomers were determined experimentally, by using the kinetic method, and theoretically, through quantum chemistry calculations. The experimental values, relative to the gas-phase acidity of phenol, are (in kJ mol−1): −1.76 ± 0.76 (2,3-Me2C6H3OH), 1.78 ± 0.29 (2,4-Me2C6H3OH), 0.83 ± 0.58 (2,5-Me2C6H3OH), −4.39 ± 0.89 (2,6-Me2C6H3OH), 5.38 ± 1.08 (3,4-Me2C6H3OH), and 1.88 ± 0.08 (3,5-Me2C6H3OH). This trend was discussed by considering the substituent effects on the thermodynamic stabilities both of the parent phenols and the corresponding phenoxide ions. The above acidity data, the literature values for 2-, 3-, and 4-methylphenol, and the substituent effects analysis allowed to develop a simple empirical method to estimate the acidity of any methyl-substituted phenol.  相似文献   

2.
李蕾  莫丹  陈大舟 《中国化学》2005,23(3):266-271
本文采用焙烧复原法研究了镁铝水滑石与吡啶二甲酸异构阴离子单体及其混合体的插层反应,实验发现镁铝水滑石对吡啶二甲酸异构阴离子存在着明显的选择性,有机酸异构体优先进入层间的顺序是:2,3-吡啶二甲酸>2,5-吡啶二甲酸>2,4-吡啶二甲酸>3,5-吡啶二甲酸>3,4-吡啶二甲酸>2,6-吡啶二甲酸。利用XRD、IR和TG测试技术对样品进行了表征,同时采用Gaussian-98软件包中ab initio 分子轨道法(HF/6-31G)计算了吡啶二甲酸异构阴离子的分子结构,理论结合实验探讨了阴离子在水滑石层间可能的空间构型,分析了其结构与插层行为的关系。研究表明镁铝水滑石层状材料插层过程中具有分子识别能力,可用于分离有机异构阴离子。  相似文献   

3.
The O?H bond dissociation energy (D O?H) has been determined for eight alkylseleno-substituted phenols, one alkyltelluro-substituted phenol, and one alkyltelluro-substituted pyridinol. D O?H has been estimated by the intersecting-parabolas method from kinetic data using five reference compounds: α-tocopherol (D O?H = 330.0 kJ/mol), 3,5-di-tert-butyl-4-methoxyphenol (D O?H = 347.6 kJ/mol), 4-methylphenol (D O?H = 361.6 kJ/mol), 2,6-di-tert-butyl-4-methylthiophenol (D O?H = 336.3 kJ/mol), and 2,6-di-ter-tbutyl-4-methylphenol (D O?H = 338.0 kJ/mol). The following D O?H values (kJ/mol) have been obtained: 335.9 for 2,5,7,8-tetramethyl-2-phytyl-6-hydroxy-3,4-dihydro-2H-1-benzoselenopyran, 342.6 for 2-methyl-5-hydroxy-2,3-dihydrobenzoselenophene, 333.5 for 2,4,6,7-tetramethyl-5-hydroxy-2,3-dihydrobenzoselenophene, 339.4 for 2-tert-butyl-4-methoxy-6-octylselenophenol, 357.9 for dodecyl 3-(4-hydroxyphenyl) propyl selenide, 348.5 for dodecyl 3-(3,5-dimethyl-4-hydroxyphenyl)propyl selenide, 350.9 for dodecyl 3-(3-tert-butyl-4-hydroxyphenyl)propyl selenide, 338.0 for dodecyl 3-(3,5-di-tert-butyl-4-hydroxyphenyl) propyl selenide, 343.0 for 2,6-di-tert-butyl-4-(tellurobutyl-4′-phenoxy)phenol, and 338.8 for 6-octyltelluro-3-pyridinol. The stabilization energies of phenoxyl radicals containing R substituents (X = O, S, Se, Te) have been compared.  相似文献   

4.
DFT calculations, at the B3LYP/TZVP level of theory for pyrocatechuic acid (2,3-dihydroxybenzoic acid, 2,3-DHBA), 2,3-dihydroxy-pyridine 2,3-DHPY and their ionized and oxidized forms, have been performed, in combination with experimental data. 1H, 13C, 2D COSY NMR, IR and electronic spectra were coupled to the theoretical calculations. The geometrical parameters were checked by reported crystallographic data. The neutral form of pyrocatechuic acid is the most stable, regarding its ionized (mono-, di- or tri-anions) and oxidized ([2,3-DHBA-sqH], [2,3-DHBA-sq]2−, [2,3-DHBA-q]) species. The most stable conformer 2,3-DHBA-H3 displays the COOH– group co-planar to the catechol ring, hydrogen bonded with OH(2). In the [2,3-DHBA-H2] the stable conformer shows the presence of protonated COOH, while OH(2) is ionized. The tri-anion is the form of 2,3-DHBA with the highest energy. Among the protonated semiquinone radical forms [2,3-DHBA-sqH], more stable is the OH(3)-oxidized, cited 21.3 kcal/mol lower in energy from the OH(2)-oxidized; in this latter the COO group lies perpendicular to the benzene ring. The same calculation procedure fitted on the oxygenated [2,3-DHBA-H-O2]2− shows a weak π-bonding between O(2) and dioxygen, strongly H–bonded to OH(3), while the C(2)–O bond order increases. The different way of 2,3-DHBA oxidation parallels the different, from 3,4-isomer, degradation products. Our DFT calculations show that the keto/enol tautomeric forms of the neutral 2,3-DHPY-H2 differ by 5.02 kcal/mol. Both species give, upon ionization, the [2,3-DHPY-H] with the OH(2) deprotonated. The electronic density distribution of [2,3-DHPY-q] justifies further reactions (degradation or dienic addition) as experimentally observed.  相似文献   

5.
The lipophilicity of the nitrophenols, expressed as a water-solvent partition coefficient, P, has been investigated using the solvation equation, log P = c + eE + sS + aA + bB + vV. It is shown that this equation accounts quantitatively for lipophilicity in a selection of water-solvent systems, viz: octanol, 1,2-dichloroethane, and cyclohexane. In the latter two systems, the major factor in the increased lipophilicity of 2-nitrophenol over 3- and 4-nitrophenol is the lack of hydrogen bond acidity of 2-nitrophenol. The water-octanol system differs in that the a coefficient is effectively zero, so that hydrogen bond acidity of solutes plays no part, and the three mononitrophenols then have similar lipophilicities. The dinitrophenols and picric acid are similarly discussed. The hydrogen bond acidity of 2,3-dinitrophenol (0.67) is very much larger than that of 2,4- or 2,5-dinitrophenol (0.09 and 0. 11), indicating a very much reduced internal hydrogen bonding. A similar but much smaller effect occurs with 2,6-dinitrophenol (A = 0. 17). Picric acid has a moderate hydrogen bond acidity (0.46) so that the phenolic OH is still available for external hydrogen bonding. These results are confirmed by ab initio calculations which show that 2,3- and 2,6-dinitrophenol and picric acid are significantly distorted away from planarity, which apparently disrupts their internal hydrogen bonding.  相似文献   

6.
Electrophilic trisubstituted ethylenes, dihalogen ring-substituted ethyl 2-cyano-3-phenyl-2-propenoates, RPhCH?C(CN)CO2C2H5 (where R is 2,3-diCl, 2,4-diCl, 2,6-diCl, 3,4-diCl, 3,5-diCl, 2,3-diF, 2,4-diF, 2,5-diF, 2,6-diF, 3,4-diF, 3,5-diF) were prepared and copolymerized with styrene. The monomers were synthesized by the piperidine catalyzed Knoevenagel condensation of ring-substituted benzaldehydes and ethyl cyanoacetate, and characterized by CHN analysis, IR, 1H and 13C-NMR. All the ethylenes were copolymerized with styrene (M1) in solution with radical initiation (ABCN) at 70°C. The composition of the copolymers was calculated from nitrogen analysis, and the structures were analyzed by IR, 1H and 13C-NMR. The order of relative reactivity (1/r 1) for the monomers is 3,4-diCl (1.89) > 2,4-diCl (1.84) > 3,5-diCl (1.40) > 2,6-diCl (1.21) > 2,4-diF (1.16) > 2,3-diF (1.01) > 2,3-diCl (0.74) > 3,4-diF (0.52) > 2,6-diF (0.45) > 3,5-diF (0.44) > 2,5-diF (0.33). Relatively high Tg of the copolymers in comparison with that of polystyrene indicates a decrease in chain mobility of the copolymer due to the high dipolar character of the trisubstituted ethylene monomer unit. Decomposition of the copolymers in nitrogen occurred in two steps, first in the 250–500°C range with residue (2.6–5.0 wt%), which then decomposed in the 500–800°C range.  相似文献   

7.
Two di-tert-butylphenols incorporating an N-methylbenzimidazole moiety in the ortho or para position have been synthesised ((Me)OH and (pMe)OH, respectively). Their X-ray structures evidence a hydrogen bond between the phenolic proton and the iminic nitrogen atom, whose nature is intra- and intermolecular, respectively. The present studies demonstrate that (Me)OH is readily oxidised by an intramolecular PET mechanism to form the hydrogen-bonded phenoxyl-N-methylbenzimidazolium system ((Me)OH)(.+) , whereas oxidation of (pMe)OH occurs by intermolecular PET, affording the neutral phenoxyl benzimidazole ((pMe)O)(.) system. The deprotonations of (Me)OH and (pMe)OH yield the corresponding phenolate species ((Me)O)(-) and ((pMe)O)(-), respectively, whilst that of the previously reported (H)OH (analogous to (Me)OH but lacking the N-methyl group) produces an unprecedented hydrogen-bonded phenol benzimidazolate species, as evidenced by its X-ray structure. The latter is believed to be in equilibrium in solution with its tautomeric phenolate form, as suggested by NMR, electrochemistry and DFT studies. The one-electron oxidations of the anions occur by a simple ET process affording phenoxyl radical species, whose electronic structure has been studied by HF-EPR spectroscopy and DFT calculations. In particular, analysis of the g(1) tensor shows the order 2.0079>2.0072>2.0069>2.0067 for ((Me)O)(.), ((H)O)(.), ((Me)OH)(.+) and ((H)OH)(.+), respectively. ((Me)O)(.) exhibits the largest g(1) tensor (2.0079), consistent with the absence of intramolecular hydrogen bond. The g(1) tensor of ((H)O)(.) is intermediate between those of ((Me)OH)(.+) and ((Me)O)(.) (g(1)=2.0072), indicating that the phenoxyl oxygen is hydrogen-bonded with a neutral benzimidazole partner.  相似文献   

8.
Graphite electrode modified with laccase from Cerrena unicolor served as a biosensor for detection of 30 phenolic compounds with different structures. Some correlations of the sensor response to the structures of substrates are discussed. This biosensor responded to: (i) nanomolar concentrations of some of the selected phenolic compounds, e.g., 2,6-dimethoxyphenol, coniferyl alcohol, caffeic acid, DOPAC and hydroquinone, (ii) micromolar concentrations, e.g., ferulic acid, syringic acid, dopamine, 3,4-dihydroxybenzoic acid and dl-noradrenaline, and (iii) millimolar concentrations in the case of phenol and 4-hydroxybenzaldehyde. Among the ortho- or para-substituted phenols, the sensitivity of the C. unicolor laccase-modified electrode increased in the following order H, CH(3), OH, OCH(3) and NH(3)(+) but in the case of para-substituted phenols, the K(m)(app) values were lower. The sensitivity of the laccase electrode increased with an additional OH group in para-substituted phenols. In the case of the selected compounds, kinetic data from electrochemical flow injection system were compared with those obtained from experiments in solution.  相似文献   

9.
Ab initio and density functional theory (DFT) calculations have been performed to understand the destruction chemistry of an important organophosphorus insecticide O,O-dimethyl O-(3-methyl-4-nitrophenyl) phosphorothioate, fenitrothion (FN), toward nucleophilic attack. Breaking of the P-OAr linkages through nucleophilic attack is considered to be the major degradation pathway for FN. One simple nucleophile, hydroxide (OH(-)), and two different α-nucleophiles, hydroperoxide (OOH(-)) and hydroxylamine anion (NH(2)O(-)), have been considered for this study. Nucleophilic attack at the two different centers, S(N)2@P and S(N)2@C, has been monitored, and the computed reaction energetics confirms that the S(N)2@P reactions are favorable over the S(N)2@C reactions for all the nucleophiles. All electronic structure calculations for the reaction are performed at DFT-B3LYP/6-31+G(d) level of theory followed by a refinement of energy at ab initio MP2/6-311++G(2d,2p) level. The effect of aqueous polarization on both the S(N)2 reactions is taken into account employing the conductor-like screening model (COSMO) as well as polarization continuum model (PCM) at B3LYP/6-31+G(d) level of theory. Relative performance of the two α-nucleophiles, OOH(-) and NH(2)O(-), at the P center has further been clarified using natural bond orbital (NBO), conceptual DFT, and atoms in molecules (AIM) approaches. The strength of the intermolecular hydrogen bonding in the transition states and topological properties of the electron density distribution for -X-H···S (X = O, N) intermolecular hydrogen bonds are the subject of NBO and AIM analysis, respectively. Our calculated reaction energetics and electronic properties suggest that the relative order of nucleophilicity for the nucleophiles is OOH(-) > NH(2)O(-) > OH(-) for the S(N)2@P, whereas for the S(N)2@C the order, which gets little altered, is NH(2)O(-) > OOH(-) > OH(-).  相似文献   

10.
Treatment of easily prepared l-substituted-4-benzylidene-2,3-dioxopyrrolidines with ammonium formate produces 1,2,4,6,7,8-hexahydro-2,6-disubstituted-8-aryldipyrrolo [3,4-b:3′,4′-e]pyridine-3,5-diones (II), usually in yields of 50 to 60%. Aromatization of the dihydropyridine ring of the hexahydroderivatives II yields corresponding 1,2,6,7-tetrahydro-2,6-disubstituted-8-aryldipyrrolo[3,4-b:3′,4′-e]pyridine-3,5-diones (III). These compounds appear to be the first to incorporate the dipyrrolo[3,4-b:3′,4′-e]pyridine ring system.  相似文献   

11.
The formidable reactivity of the oxygen-evolving center near photosystem II is largely based on its protein environment that stabilizes it during catalysis. Inspired by this concept, the water-soluble Mn12 clusters Mn12O12(O2CC6H3(OH)2)16(H2O)4 (3,5DHMn12) and Mn12O12(O2CC6H3(OH)3)16(H2O)4 (3,4,5THMn12) were developed as efficient electrocatalysts for water oxidation. In this work, the role of the −OH groups in the electrocatalytic process was explored by describing the structural and electrocatalytic properties of two new Mn12 clusters, 3,4DHMn12 and 2,3DHMn12 , having one −OH group in the meta position relative to the benzoate-Mn moiety, and one at the para or ortho position, respectively. The Mn centers in 3,4DHMn12 were discovered to have lower oxidation potential compared with those in 2,3DHMn12 , and thus, 3,4DHMn12 can catalyze water oxidation with higher rate and TON than 2,3DHMn12 . Hence, the role of the −OH groups in the electrocatalysis was established, being involved in electronic stabilization of the Mn centers or in proton shuttling.  相似文献   

12.
Two supramolecular complexes: [Co(2,6-PDC)(Hdmpz)3]·H2O (1) and [Zn(2,6-PDC)(Hdmpz)2] (2) {2,6-PDC=pyridine-2,6-dicarboxylic acid, Hdmpz=3,5-dimethylpyrazol}, self-assembles via O-H?O and N-H?O hydrogen bondings into supramolecular networks, which are characterized by elemental analyses, IR spectra and single crystal X-ray diffraction analysis. Both of them consist of two-dimensional networks that are stacked together by typical hydrogen bonding interactions (i.e. O-H?O and N-H?O), which often play important roles in the formation of low-dimensional into high-dimensional supramolecular networks. In addition, quantum chemistry calculations and surface photovoltage spectroscopy are performed firstly with the complexes.  相似文献   

13.
Tetra(meso-aryl)porphyrins containing 2,6-dimethoxyphenol unit distant from the macrocycle and adjacent to it have been prepared. Radical scavenging activity of the obtained compounds has been estimated in the model reaction of ethylbenzene oxidation initiated by azobisisobutyronitrile. It has been shown that the phenol hydroxy group ability to inhibit ethylbenzene oxidation depends on the environment of phenol group.  相似文献   

14.
We have carried out extensive calculations for neutral, cationic protonated, anionic deprotonated phenol dimers. The structures and energetics of this system are determined by the delicate competition between H-bonding, H-π interaction and π-π interaction. Thus, the structures, binding energies and frequencies of the dimers are studied by using a variety of functionals of density functional theory (DFT) and M?ller-Plesset second order perturbation theory (MP2) with medium and extended basis sets. The binding energies are compared with those of highly reliable coupled cluster theory with single, double, and perturbative triple excitations (CCSD(T)) at the complete basis set (CBS) limit. The neutral phenol dimer is unique in the sense that its experimental rotational constants have been measured. The geometry of the neutral phenol dimer is governed by the hydrogen bond formed by two hydroxyl groups and the H-π interaction between two aromatic rings, while the structure of the protonated/deprotonated phenol dimers is additionally governed by the electrostatic and induction effects due to the short strong hydrogen bond (SSHB) and the charges populated in the aromatic rings in the ionic systems. Our salient finding is the substantial differences in structure between neutral, protonated, and deprotonated phenol dimers. This is because the neutral dimer involves in both H(π)···O and H(π)···π interactions, the protonated dimer involves in H(π)···π interactions, and the deprotonated dimer involves in a strong H(π)···O interaction. It is important to compare the reliability of diverse computational approaches employed in quantum chemistry on the basis of the calculational results of this system. MP2 calculations using a small cc-pVDZ basis set give reasonable structures, but those using extended basis sets predict wrong π-stacked structures due to the overestimation of the dispersion energies of the π-π interactions. A few new DFT functionals with the empirical dispersion give reliable results consistent with the CCSD(T)/CBS results. The binding energies of the neutral, cationic protonated, and anionic deprotonated phenol dimers are estimated to be more than 28.5, 118.2, and 118.3 kJ mol(-1), respectively. The energy components of the intermolecular interactions for the neutral, protonated and deprotonated dimers are analyzed.  相似文献   

15.
运用密度泛函理论(DFT),采用Mo16S32团簇模型,在PW91/DNP水平上研究了噻吩(TP)及一系列烷基噻吩类硫化物如2-甲基噻吩(2-MT)、3-甲基噻吩(3-MT)、2,3-二甲基噻吩(2,3-DMT)、2,4-二甲基噻吩(2,4-DMT)、2,5-二甲基噻吩(2,5-DMT)及3,4-二甲基噻吩(3,4-DMT)等在加氢脱硫催化剂MoS2上的吸附行为.结果表明,在η1S吸附构型中,Mo16S32团簇对烷基噻吩吸附能力的顺序为2,5-DMT>2,4-DMT≈2,3-DMT>2-MT>3,4-DMT>3-MT>TP.通过键长、Mayer键级、Mulliken电荷分析可知,当噻吩环的2-或5-位不含甲基时,吸附能随硫原子电荷密度的增加而增大;2-或5-位含甲基时,甲基与团簇上相邻的Mo原子发生了弱的相互作用,使吸附能增大;虽然2,5-DMT的2-和5-位均含有甲基,但甲基离团簇上相邻的Mo较远,相互作用较小,吸附能较2,3-DMT和2,4-DMT增加的较少.文中还对各硫化物在MoS2催化剂上的加氢脱硫反应进行了讨论.  相似文献   

16.
Diethyl 4‐(2,5‐di­methoxy­phenyl)‐2,6‐di­methyl‐1,4‐di­hydro­pyridine‐3,5‐di­carboxyl­ate, C21H27NO6, (I), diethyl 4‐(3,4‐di­methoxy­phenyl)‐2,6‐di­methyl‐1,4‐di­hydro­pyridine‐3,5‐di­carboxyl­ate, C21H27NO6, (II), and diethyl 2,6‐di­methyl‐4‐(3,4,5‐tri­methoxy­phenyl)‐1,4‐di­hydro­pyridine‐3,5‐di­carboxyl­ate, C22H29NO7, (III), crystallize with hydrogen‐bonding networks involving the H atom bonded to the N atom of the 1,4‐di­hydro­pyridine ring and carbonyl O atoms in (I) and (II). Unusually, (III) shows O atoms of methoxy groups serving as hydrogen‐bond acceptors.  相似文献   

17.
Abstract

2,4-; 2,6-; 2,3-; 3,4-; 2,5-; and 3,5-dichlorophenols by reaction with chlorosulfonic acid were converted to the following substituted benzenesulfonyl chlorides: 3,5-dichloro-2-hydroxy-; 3,5-dichloro-4-hydroxy-; 2,3-dichloro-4-hydroxy-; 4,5-dichloro-2-hydroxy-; 2,5-dichloro-4-hydroxy-; and 2,6-dichloro-4-hydroxy-respectively. In addition o-chlorophenol gave 5-chloro-4-hydroxybenzene-1,3-bis-sulfonyl chloride. The various sulfonyl chlorides have been condensed with nucleophilic reagents, e.g. ammonia, amines, hydrazine, phenylhydrazine, N, N-dimethylhydrazine, and sodium azide. 3,5-Dichloro-2-hydroxybenzenesulfonyl azide has been reacted with norbornene, triphenylphosphine, dimethylsulfoxide, and cyclohexene. 3,5-Dichloro-2-hydroxybenzenesulfonyl chloride with phenylisocyanate gave the 2-(N-phenyl-carbamoyloxy) derivative which on heating gave a heterocyclic compound. The chlorohydroxybenzenesulfonyl derivatives are of interest as potential herbicides and their ir and nmr spectral characteristics are briefly discussed.  相似文献   

18.
A new aromatic glycoside, shamiminol was isolated from the stem bark of Bombax ceiba along with the known constituents stigmasta-3,5-diene, lupenone, (+/-)-lyoniresinol 2a-O-beta-D-glucopyranoside and opuntiol, obtained for the first time from this plant. The structure of shamiminol was elucidated on the basis of extensive 1D- and 2D-NMR spectroscopic and mass spectrometric studies as 3,4,5-trimethoxyphenol 1-O-beta-D-xylopyranosyl-(1 --> 2)-beta-D-glucopyranoside (1).  相似文献   

19.
为了研究模板分子中作用基团的数目和位置对印迹聚合物印迹效应的影响, 分别以含有羟基数目和位置不同的羟基苯甲酸化合物3,4,5-三羟基苯甲酸(3,4,5-THBA), 3,4-二羟基苯甲酸(3,4-DHBA), 2,4-二羟基苯甲酸(2,4-DHBA)和3-羟基苯甲酸(3-HBA)为模板分子, 以丙烯酰胺为功能单体, 乙二醇二甲基丙烯酸酯为交联剂和乙腈(MeCN)为致孔剂, 采用非共价本体聚合方法制备了对应的印迹聚合物, 用色谱法评价了其分子识别性能. 结果表明, 制备的印迹聚合物对相应的模板分子均具有印迹效应, 在流动相H2O/MeCN(体积比1/99)中, 各印迹聚合物对相应的模板分子3,4,5-THBA, 3,4-DHBA, 2,4-DHBA和3-HBA的印迹因子分别为5.51, 5.55, 2.60和2.03. 通过与同样条件下制备的龙胆酸(GA)、水杨酸(SA)和对-羟基苯甲酸(4-HBA)印迹聚合物对其模板分子印迹效应的比较发现, 模板分子中作用基团数目越多, 印迹效率越高; 模板分子中作用基团-COOH和-OH的相对位置对印迹效率影响很大, 当-COOH和-OH在苯环上处于对位时的印迹效率, 高于其处于间位的印迹效率; 当-COOH和-OH在苯环上处于邻位时, 由于形成分子内氢键会降低其印迹效率. 实验还发现, 3,4-DHBA的印迹聚合物可以实现其结构类似物3,4,5-THBA和2,4-DHBA的基线分离, 为生物活性组分3,4,5-THBA的分离和测定提供了依据.  相似文献   

20.
Mo(PMe3)6 reacts with 2,6-Ph2C6H3OH to give the eta 6-arene complex [eta 6-C6H5C6H3(Ph)OH]Mo(PMe3)3 which exhibits a non-classical Mo...H-OAr hydrogen bond; DFT calculations indicate that the hydrogen bonding interaction facilitates oxidative addition of the O-H bond to give [eta 6,eta 1-C6H5C6H3(Ph)O]Mo(PMe3)2H.  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