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 203 毫秒
1.
The N—H and O—H bond dissociation energies in 4-hydroxydiphenylamine Ph—NH—C6H4—OH (D NH= 353.4, D OH=339.3 kJ mol–1) and its semiquinone radicals D NH(Ph—NH—C6H4—O·) = 273.6, D OH(Ph—N·—C6H4—OH) = 259.5 kJ mol–1 were first estimated using the parabolic model and experimental data (rate constants) on two elementary reactions with participation of N-phenyl-1,4-benzoquinonemonoimine (2). One of the reactions, namely, that of 2 with aromatic amines, was studied in this work using a specially developed method.  相似文献   

2.
Electrolysis of the system Ti(IV)–NH2OH–C6H6 in an 11 M H2SO4 solution shows that using an organic solvent (acetic acid, acetonitrile) during cathodically initiated amination of aromatic substrates permits the production of aromatic amines with the overall yield by hydroxylamine reaching 91%. Due to a chain mechanism of radical substitution, the benzene amination in electrolytes containing 5 M CH3COOH and 5.5 M CH3CN terminates largely upon consuming 70–75 and 50–55% of the charge required theoretically for a one-electron process. The maximum efficiency of electrochemical amination is observed at low hydroxylamine conversions and the overall current efficiency for mono- and disubstituted products of the benzene amination may exceed in these conditions 750%.  相似文献   

3.
Electrochemical amination of benzene in sulfuric acid electrolytes is studied and experimental conditions for highly efficient synthesis of primary aromatic amino compounds are determined. In the electrolysis of Ti(IV)–NH2OH–C6H6 in 11 M H2SO4 solutions containing acetic acid or acetonitrile as organic solvents, aniline and isomeric phenylenediamines are obtained with the total yields by hydroxylamine of 95.6 and 99.6%, respectively. A monoamino compound is the main product of radical substitution in acidic organo-aqueous media. It is found that the use of acetonitrile in electrochemical process is limited to certain sulfuric acid concentrations and temperatures.  相似文献   

4.
A linear relationship was found between the first reduction potentials (E°red) and electron affinities (EA) for fullerenes C60 and C70, their hydro- and fluoro-derivatives, and aromatic hydrocarbons: E°red = –3.04 + 0.81·EA. This equation was used to estimate the unknown values of EA = 2.45 eV for C60H2, 2.47 eV for C70H2, –0.15 eV for C70H36—38, –0.41 eV for C70H44—46, and E°red = –1.74—–1.91 V (vs. Fc0/+) for C60H18.  相似文献   

5.
The molecular and electronic structures of closo-hexaboranes B6H6 2–, B6H7 , and B6H8 and closo-heterohexaboranes XYB4H4 (X = Y = CH, N; X = BH, Y = CH, N, NH, O) were studed by the ab initio (MP2(full)/6-311+G**) and density functional (B3LYP/6-311+G**) methods. The bridging H atoms in closo-hexaboranes B6H7 and B6H8 can undergo facile low-barrier migrations around the boron cage (the barrier heights are about 10—15 kcal mol–1). All heteroboranes having octahedron-like structures with hypercoordinated N and O atoms are rather stable and can be the subject of synthetic research efforts.  相似文献   

6.
The solid—liquid equilibria of the ternary system H2O—Al(NO3)3—Mg(NO3)2 were studied at –30, –20, –10 and 0°C by using a synthetic method which allows to detemine all the characteristic points of isothermal sections. The stable solid phases which appear are respectively: ice, Al(NO3)3·9H2O, Mg(NO3)2·9H2O and Mg(NO3)2·6H2O. Neither double salts nor mixed crystals are observed in the temperature and composition field studied. Polytherm diagram layout show two invariant transformations correspond with an eutectic point and a peritectic point.This revised version was published online in November 2005 with corrections to the Cover Date.  相似文献   

7.
Reactions of an amino derivative of the closo-decaborate anion [1-B10H9NH3] with aromatic aldehydes afforded Schiff bases [1-B10H9NH=CHAr] (Ar=Ph, C6H4-2-OMe, or C6H4-4-NHCOMe). The reduction of the latter with sodium borohydride gave the corresponding benzylamino derivatives [1-B10H9NH2CH2Ar].Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 9, pp. 2004–2007, September, 2004.  相似文献   

8.
The reaction of cyclic trimeric perfluoro-o-phenylenemercury (o-C6F4Hg)3 (1) with the polyhedral [B12H11SCN]2– anion in THF at 20 °C affords the {[(o-C6F4Hg)3](B12H11SCN)}2– (4) and {[(o-C6F4Hg)3]2(B12H11SCN)}2– (5) complexes. Complex 5 was isolated as the tetrabutylammonium salt. X-ray diffraction analysis showed that this complex has a bent-sandwich structure in which the [B12H11SCN]2– anion is located between the planes of two molecules 1 and is coordinated to both these molecules through B—H—Hg bridges and S—Hg bonds. The stability constants of complexes 4 and 5 in THF (20 °C), which were determined from the IR spectroscopic data, are 16 L mol–1 and 992 L2 mol–2, respectively.  相似文献   

9.
The following substances have been isolated from an acetone extract ofFerula gigantea B. Fedtsch.: a coumarin — umbelliferone, C9H6O2, mp 230–233°C; and sesquiterpene lactones — talassin A, C25H30O7, mp 188–191°C; malaphyllinin, C24H24O7, mp 231–235°C; malaphyll, C29H32O9, mp 212–213°C; and malaphyllin, C26H28O9, mp 216–218°C. Structures have been proposed for three new sesquiterpene lactones on the basis of an analysis of their spectral characteristics.All-Union Scientific-Research Institute of Medicinal Plants, Moscow. M. V. Lomonosov Moscow State University, Botanical Garden, Moscow. Translated from Khimiya Prirodnykh Soedinenii, No. 4, pp. 490–495, July–August, 1979.  相似文献   

10.
A mononuclear iron(II) complex, Et4N[Fe(C10H6NO2)3], coordinated by three 1‐nitroso‐2‐naphtholate ligands in a fac‐N3O3 geometry, was initiated to catalyze the direct hydroxylation of aromatic compounds to phenols in the presence of H2O2 under mild conditions. Various reaction parameters, including the catalyst dosage, temperature, mole ratio of H2O2 to benzene, reaction time and solvents which could affect the hydroxylation activity of the catalyst, were investigated systematically for benzene hydroxylation to obtain ideal benzene conversion and high phenol distribution. Under the optimum conditions, the benzene conversion was 10.2% and only phenol was detected. The catalyst was also found to be active towards hydroxylation of other aromatic compounds with high substrate conversions. The hydroxyl radical formed due to the reaction of the catalyst and H2O2 was determined to be the crucial active intermediate in the hydroxylation. A rational pathway for the formation of the hydroxyl radical was proposed and justified by the density functional theory calculations. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

11.
Reduction of the binuclear PdII complexes Pd2(OCOR)2(o-CH2C6H4—NO)2 (1) and Pd2(OCOR)2(o-PhN—C6H4—NO)2 (2) (where R = Me, CF3, But, or Ph) by sodium borohydride, an ethanolic solution of KOH, or molecular hydrogen was examined. The first stage of reduction was demonstrated to afford metallic palladium and aromatic amines, viz., o-toluidine o-Me—C6H4—NH2 from complex 1 and aniline Ph—NH2 from complex 2. The reactions with molecular hydrogen involve deeper stages to yield cyclic ketones (o-methylcyclohexanone and cyclohexanone) and then cycloalkanes (methylcyclohexane and cyclohexane, respectively). The latter reactions are accompanied by elimination of N2. The mechanism of reduction of complexes 1 and 2 with molecular hydrogen was proposed.  相似文献   

12.
Alkanes and cycloalkanes (isobutane, butane, isopentane, isohexane, and methylcyclopentane) react with benzene or bromobenzene at 0–20 °C in the presence of RCO+Al2X7 complexes (R=Me, Pr, or Ph; X=Cl or Br) to give products of the alkylacylation of arenes. The yields of alkylated aromatic ketones reach 60–87 % in 5–30 min, whereas the yields of unalkylated aromatic ketones (the competitive reaction) reach 0–40 %. The reactions of isobutane or isopentane with benzene result exclusively inpara isomers oft-BuC6H4COR or a mixture of Me2(Et)CC6H4COR and Me(i-Pr)CHC6H4COR isomers (11), respectively. The reaction of isobutane with benzene also proceeds regioselectively and gives only one isomer, 2-Br-t-BuC6H4COR.For Part 2 seeIzv. Akad. Nauk, Ser. Khim., 1991, No 1, 105 [Bull. Acad. Sci. USSR. Div. Chem. Sci., 1991, No 1, 90 (Engl. Transl)].Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 7, pp. 1253–1257, July, 1993.The authors express their gratitude to B. I. Bakhmutov for his assistance in interpreting the spectra.  相似文献   

13.
Zusammenfassung Salicylsäurechlorid (1) reagiert mit aromatischen Thioamiden unter HCl- und H2S-Abspaltung zu 4H-1,3-Benzoxazinonen (2–5), welche mit verd. HCl zu N-Acylsalicylsäureamiden (6–9) gespalten werden.
Salicyloyl chloride (1) reacts with aromatic thioamides to 4H-1.3-benzoxazinones (2–5), which can be hydrolized with dil. HCl yielding N-acyl-salicyloyl-amides (6–9).
  相似文献   

14.
Zusammenfassung Die Umsetzung von Diphenylphosphinigsäurechlorid mit Alkalisalzen aromatischer Sulfinsäuren führt unter Reduktion am Schwefel, Oxydation am Phosphor und Ausbildung einer Bindung zwischen Phosphor und Schwefel, zu Diphenylthiophosphinsäure-S-arylestern: (C6H5)2P(O)–S–Ar (Ar=C6H5, 4-CH3C6H4, 4-ClC6H4, 2-Cl-5-CH3-C6H3, 2-C10H7). Die besten Ausbeuten (40–85%) wurden mitDMF als Lösungsmittel erhalten. Aliphatische Phosphinigsäurechloride geben keine Bindung zwischen P und S. Ebenso tritt keine Umsetzung ein, wenn im Säurechlorid eine P–O-Struktur vorliegt. Auch die Umsetzung von (C6H5)2PCl und (C6H5)2P(O)Cl mit Sulfonsäuren anstelle der Sulfinsäuren führt zu keiner P–S-Verknüpfung. Diese Phosphor-Schwefel-Bindung in den Thiophosphinsäure-S-estern stellt die schwächste Stelle im Molekül dar, da ein hydrolytischer, oxydativer oder reduktiver Angriff diese Bindung wieder löst.
Reaction of diphenylchlorophosphine with alkali salts of aromatic sulfinic acids leads to diphenylthiophosphinates; reduction occurs at the sulfur atom, oxidation at the phosphorus atom and a bond between phosphorus and sulfur is formed: (C6H5)2P(O)–S–Ar, Ar=C6H5, 4-CH3C6H4, 2-ClC6H4, 2-Cl-5-CH3-C6H3, 2-C10H7. The best yields were obtained in dimethyl formamide as solvent. With aliphatic chlorophosphines no bond formation beetween sulfur and phosphorus occurred. Similarly no reaction was observed, when a phosphorus atom in a higher state of oxidation was present, as for example in diphenylphosphonylchloride. Also, no reaction took place when sulfonic instead of sulfinic acids were used. The weakest bond found to exist in the diphenylthiophosphinates was the P–S-linkage, which readily undergoes hydrolytic, oxidative or reductive cleavage.
  相似文献   

15.
A novel O—N—N—O-type tetradentate ligand H4mda (H4mda = malamido-N,N-diacetic acid) and the corresponding square-planar copper(II) complexes have been prepared and characterized. The mda4– ligand coordinates to the copper(II) ion via two pairs of deprotonated ligating atoms (two carboxylate oxygens and two deprotonated amide nitrogens) with in-plane square chelation. A four-coordinate, square-planar geometry has been established crystallographically for the [Co(H2O)6][Cu(mda)] · 2H2O complex. Structural data correlating the square-planar geometry of the [Cu(mda)]2– unit are discussed in relation to information obtained for similar complexes. The i.r., electronic, absorption and reflectance spectra of the complexes are analysed in comparison with related complexes of known geometries.  相似文献   

16.
ansa-Metallocenes (5:5-C24H16)M(THF)2 (M = Sm (1), Yb (2), Ca (3)) and (5:5-C24H16)MI(THF) (M = Dy (8), Er (9), Tm (10), Lu (11)) were prepared in 50—90% yields by the in situ reactions of two equivalents of potassium acenaphthylenide K+C12H8 with MI2 or MI3, respectively. Complexes 2 and 3 were also obtained by direct reduction of acenaphthylene with ytterbium and calcium naphthalenides, respectively. An ESR signal of the acenaphthylene radical anion, which was observed upon dissolution of compound 2 in THF, indicates that the [C24H16]2– ansa-ligand dissociated into two [C12H8]·– radical anions. Hydrolysis of complex 2 in benzene afforded 1,1",3,3"-tetrahydro-3,3"-biacenaphthylene (4) and 3,3",4,4"-tetrahydro-3,3"-biacenaphthylene (5). The reaction of complex 2 with ZrCl4 and the reaction of compound 3 with Me3SiCl proceeded with the cleavage of the C—C bond between two acenaphthylene fragments of the [C24H16]2– ansa-ligand to produce (2-C12H8)ZrCl2(THF)3 (6) and bis(trimethylsilyl)acenaphthene (Me3Si)2C12H8 (7), respectively. Compounds 1—3, 6, 7, and 11 were characterized by 1H and 13C NMR spectroscopy. The temperature dependence of the 1H NMR spectrum of compound 11 in tetrahydrofuran is indicative of the dynamic exchange of the solvent molecules in the coordination sphere of the Lu atom. After cooling of the solution to 210 K, the dynamic process was terminated as evidenced by the nonequivalence of the 1H signals of two acenaphthylene fragments. According to the X-ray diffraction data for complex 11, dimerization of two acenaphthylene radical anions at the Lu atom gave rise to the rac-ansa-metallocene structure. In compound 11, the Lu atom is 5-coordinated by two five-membered rings of the acenaphthylene ligands and also by the I atom and the THF molecule. The coordination environment about the Lu atom is a distorted tetrahedron. The average distance between the lutetium atom and the carbon atoms of the five-membered rings is 2.623 .  相似文献   

17.
Xiong  Ya  He  Chun  An  Tai-Cheng  Cha  Chang-Hong  Zhu  Xi-Hai  Jiang  Shaoji 《Transition Metal Chemistry》2003,28(1):69-73
In the neutral title complex [Cu(C4N2H3)2(H2O)3] or [Cu(BBR)2(H2O)3] (BBR = Barbiturate), the CuII ion, in the slightly distorted square-pyramidal geometry, is coordinated by two O atoms of the two monodentate barbiturates and three O atoms of three water ligands. The average bond length of Cu—O (BBR) is 1.981(5) Å and the average bond length of Cu—O (H2O) at the basal sites is 1.94(5) Å, i.e. much shorter than that of Cu—O (H2O) [2.175(11) Å]. The crystal structure is characterized by an extensive network of hydrogen bonds in which each [Cu(BBR)2(H2O)3] entity links to six adjacent [Cu(BBR)2(H2O)3] by O(C=O) ··· H—O(H2O) bonds. Tautomerism in the coordination process for BBR was found from the crystal structure and i.r. spectral analysis. The interaction of CuII and BBR in aqueous solution was also investigated by electronic spectra and electrochemical method. It was observed that the copper ion could not only form the [Cu(BBR)2(H2O)3] complex in aqueous but also catalyze the decomposition of BBR at pH 1.1.  相似文献   

18.
Complexation in the Co(II)–H6X–H2O, Ni(II)–H6X–H2O, and Co(II)–Ni(II)–H6X–H2O systems (H6X is nitrilotrimethylenephosphonic acid) was studied by spectrophotometry. The formation of binuclear complexonates Ni2H2X · 7H2O, Co2H2X · 5H2O, and NiCoH2X · 6H2O was demonstrated. These compounds were isolated from the solution, their composition was determined, the thermal stability was studied, and the kinetic parameters of dehydration were calculated.  相似文献   

19.
Solubility properties of the ternary systems of Cr(NO3)3–His–H2O, Cr(NO3)3–Met–H2O, and CrCl3–His–H2O (His—histidine, Met—methionine) have been investigated in the whole concentration range by the phase equilibrium semimicromethod, and the corresponding phase diagrams have been constructed. It was shown that the new complexes Cr(His)(NO3)3 · 3H2O, Cr(His)2(NO3)3 · 3H2O, Cr(His)3(NO3)3 · 3H2O, Cr(His)Cl3 · H2O, Cr(His)2Cl3 · H2O, and Cr(His)3Cl3 · H2O are formed in the Cr(NO3)3/CrCl3–His–H2O system, while Cr(Met)(NO3)3 · H2O and Cr(Met)2(NO3)3 · H2O complexes are formed in the system Cr(NO3)3–Met–H2O. Under the guidance of the phase diagrams, the complexes were prepared and characterized by chemical and elemental analysis, IR spectroscopy, and thermogravimetry data. The influences of metal cations, anions and the structures of amino acids on the formation of complexes were discussed.  相似文献   

20.
Yuan  Ai-Hua  Lu  Lu-De  Shen  Xiao-Ping  Chen  Li-Zhuang  Yu  Kai-Bei 《Transition Metal Chemistry》2003,28(2):163-167
A cyanide-bridged FeIII–FeII mixed-valence assembly, [FeIII(salen)]2[FeII(CN)5NO] [salen = N,N-ethylenebis(salicylideneiminato)dianion], prepared by slow diffusion of an aqueous solution of Na2[Fe(CN)5NO] · 2H2O and a MeOH solution of [Fe(salen)NO3] in an H tube, has been characterized by X-ray structure analysis, i.r. spectra and magnetic measurements. The product assumes a two-dimensional network structure consisting of pillow-like octanuclear [—FeII—CN—FeIII—NC—]4 units with dimensions: FeII—C = 1.942(7) Å, C—N = 1.139(9) Å, FeIII—N = 2.173(6) Å, FeII—C—N = 178.0(6)°, FeIII—N—C = 163.4(6)°. The FeII—N—O bond angle is linear (180.0°). The variable temperature magnetic susceptibility, measured in the 4.8–300 K range, indicates the presence of a weak intralayer antiferromagnetic interaction and gives an FeIII–FeIII exchange integral of –0.033 cm–1.  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