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 187 毫秒
1.
The standard Gibbs free energies of transfer from pyridine to pyridine+diluent mixtures are reported for the ZnCl2py2 complex at 25°C. Their variations with varying mixed solvent composition agree with those expected for regular solutions when the diluents are polar and basic while marked deviations are observed for non-polar, inert diluents. For protic diluents the free energies of transfer of ZnCl2py2 exhibit maxima, indicating specific interaction of the complex with both solvent components in these systems. The free energies of transfer of ZnCl2py2 to the pure diluents plotted vs. the solubility parameter of the latter reveal separate Hildebrand correlations for inert and for basic solvents, with the free energies of transfer of the complex to the protic solvents deviating from both. Weak specific interactions of ZnCl2py2 with the basic solvents and stronger interactions with the protic solvents are inferred and further examined in the light of the free energy of transfer data for the related tetrahedral complexes ZnBr2py2, ZnCl2(α-pic)2, CoCl2(α-pic)2, and ZnBr2(α-pic)2.1H-NMR spectrum of the pyridine ligands of ZnCl2py2 confirm their involvement in weak hydrogen-bond formation with the basic solvents, while the downfield shift observed for the solvent chloroform is consistent with hydrogen bond formation of the latter with the chloro-ligands.  相似文献   

2.
In dissociation experiments of H2O2 under shock wave conditions, the spectra of H2O2 and HO2 have been observed in the UV at 2200 ≤ 2800 Å. By the use of these spectra the H2O2 decomposition in the presence of H2 and CO at 870 ≤ T ≤ 1000°K has been analyzed. It was found that in this temperature range, in contrast to low temperature behavior, reactions of H atoms with H2O2 and with HO2 are equally important. The rate of the reaction H + H2O2 ← HO2 + H2 was estimated in comparison with the rate of the reaction between H and HO2. Good agreement between calculated and measured concentration profiles of HO2 and H2O2 was obtained.  相似文献   

3.
The interaction of Ph3PPD(OAc)22 with molecular H2 yields a binuclear complex of zero-valent palladium, (Ph3P)2Pd2. This complex interacts reversibly with H2 in CH2Cl2, yielding (Ph3P)2Pd2H2. In argon atmosphere (Ph3P)2Pd2 reacts with [Ph3PPd(OAc)22 to form a binuclear complex of PdI with a metal—metal bond. These data, as well as the results of kinetic studies of the reactions between [Ph3PPd(OAc)22 and H2, are in agreement with an autocatalytic mechanism for the process, including catalysis of the reduction of PdII complexes by the Pd0 compounds. It has been established that the synthesized compound of PdII, PdI and Pd0 with the ratio P/Pd?1, are inactive in the hydrogenation of unsaturated compounds. The catalytically active complex (PPh)2Pd5 is formed when palladium acetate reacts with (Ph3P)2Pd2 in the presence of H2. The same compound is formed when a solution of (Ph3P)2Pd2 is treated with a mixture of H2 and O2 (or H2O2 in an atmosphere of H2). (PPh)2Pd5 is an effective catalyst for the hydrogenation of olefins, dienes, acetylenes, aldehydes, organic peroxides, quinones, O2, Schiff bases, and nitro, nitroso, and azo compounds.  相似文献   

4.
The first acetylene complex of hafnium, Cp2Hf[Me3SiC=CHf(H)Cp2], was synthesized by the reaction of hafnocene dihydride Cp2HfH2 with bis(trimethylsilyl)acetylene in benzene. The reaction is accompanied by elimination of the Me3Si group from the molecule of the initial acetylene, as a result of which the acetylenide derivative of hafnium Cp2Hf(C=CSiMe3)(H) acts as an acetylene ligand in the complex. Under analogous conditions, the reaction of zirconocene dihydride Cp2ZrH2 with bis(trimethylsilyl)acetylene affords an analogous acetylene complex of zirconium Cp2Zr(M3SiC=CZr(H)Cp2]. Reactions of Cp2HfH2 with tolane and 3-hexyne proceed differently than the reaction with bis(trimethylsilyl)acetylene. Here the corresponding hafnacyclopentadiene metallacycles are the final products. For preliminary communication, see Ref. 3. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 4, pp. 853–856, April, 1997.  相似文献   

5.
Complexation of ketoconazole (KET), a broad-spectrum antifungal drug, with β- and γ-cyclodextrins (CDs), heptakis (2,6-di-O-methyl)-β-CD (2,6-DM-β-CD), heptakis (2,3,6-tri-O-methyl)-β-CD (TM-β-CD), 2-hydroxypropyl-β-CD (2HP-β-CD) and carboxymethyl-β-CD (CM-β-CD) was studied. The stability constants were determined by the solubility method at pH = 6 and for 2,6-DM-β-CD and CM-β-CD at pH = 5. At pH = 6, the stability constants increased in the order: TM-β-D < γ-CD < 2HP-β-CD < β-CD < CM-β-CD < 2,6-DM-β-CD. At pH = 5, due to the increased ionization of KET, the stability constant with CM-β-CD increased and with 2,6-DM-β-CD decreased. For complexes of KET with 2HP-β-CD and 2,6-DM-β-CD, the thermodynamic parameters of complexation were determined from the temperature dependence of the corresponding stability constants. For β–γ and TM-β-CD complexes, calculations using HyperChem 6 software by the Amber force field were carried out to gain some insight into the host–guest geometry.  相似文献   

6.
Catalytic dehydration of 2‐propanol and that of 1‐butanol were performed at atmospheric pressure and 150–300°C over ZrO2 and sulfated ZrO2 (S/ZrO2) in a fixed‐bed, tubular reactor. The catalysts were characterized with XRD, elemental analysis, FT‐IR, N2 physisorption, TG/DTA, TPD, and TPR. The main structures of ZrO2 and S/ZrO2 were monoclinic and tetragonal, respectively. As ZrO2 was modified with sulfuric acid, its surface area and acid amount were greatly increased, whereas the pore volume, the pore diameter, and the particle size were reduced. Both samples owned weak basicity. For both reactions, only dehydration products of alkene and ether were obtained. The alcohol conversion enhanced remarkably with the catalyst acid amount and the surface area as well as the reaction temperature. In addition, the ether selectivity on S/ZrO2 decreased with raising the reaction temperature. The activation energy was 81.0 kJ/mol in the propene formation from 2‐propanol over S/ZrO2. The corresponding value was 94.4 kJ/mol for the dehydration of 1‐butanol.  相似文献   

7.
The [1+1] condensation of isonitrosoacetylacetone (Hisoacac) with o-phenylenediamine produces the diazepine (HLBD) (1), which reacts with Ni(OAc)2· 4H2O (1:1 molar ratio) to produce the mixed ligand complex (LBDN)Ni(OAc) (2); where LBDN is the anion of the half unit obtained by hydrolysis of one HLBD imine linkage. The reaction of (2) (1 mol) with mono-, bi- and trichloroethanoic acid (1mol) or picric acid (1mol) led to the exchange of the acetate in (2) with the anion of the added acid [(3)–(6), respectively]. The supramolecular structure of (2)–(6) is achieved through the dimerization of these complexes via intermolecular hydrogen bonding of the LBDN –NH2 group of one molecule and the monodentate acetate group of another molecule. The template reaction of o-phen with Hisoacac in the presence of Ni(OAc)2·4H2O (1:2:2 and 1:2:1 molar ratios, respectively) led to the formation of (LBDN)Ni(OAc)2Ni(isoacac) (7) and (isophen)Ni (8), respectively; H2isophen is a symmetrical Schiff base ligand formed by the (2:1) in situ condensation of Hisoacac with o-phen. The (1:1) condensation of Hisoacac with p-phen produced the half unit Hisopphen (9), whose 1:1 molar ratio reaction with Ni(OAc)2·4H2O led to the formation of (isopphen)Ni(OAc)·2H2O (10). The amino group of the isopphen ligand is available for further coordination with the nickel(II) ion to produce the metallosupramolecular complexes {[two molecules of complex (10)] [Ni(OAc)2]} and {[complex (10)] [Ni(OAc)2·H2O]} from the 2:1 and 1:1 molar ratio reactions, respectively, of (10) with Ni(OAc)2·4H2O. The 1:1 molar ratio reaction of (10) with Hisoacac led to replacement of OAc by isoacac. The suggested structures of the ligands and their coordination compounds are based on analytical, chemical, spectral data and magnetic moments.  相似文献   

8.
Reaction of 2-phenylethynyl N-tosylanilide prepared by Pd-free procedure with ZnBr2 (3 equiv) in refluxing toluene gave N-tosyl-2-phenylindole in 93% yield. Treatment of 2-phenylethynylaniline with ZnBr2 (1 equiv) in refluxing toluene resulted in the formation of 2-phenylindole in 91% yield. Catalytic ZnBr2 (0.05 equiv) effectively reacted with 2-alkynylanilines to afford 2-substituted indoles in high yields. Thus, complete Pd-free zinc catalyzed hydroamination of 2-alkynylanilines was achieved.  相似文献   

9.
A series of nitroimidazoles were subjected to hydroxymethylations under a variety of conditions. Hydroxymethylation of 1-(2-hydroxyethyl), 1-(2-acetoxyethyl), and 1-(2-chloroethyl) substituted 5-nitroimidazoles with paraformaldehyde in dimethyl sulfoxide yielded the respective 2-hydroxymethyl analogs (5–7). However, attempts to hydroxymethylate 1-(2-hydroxyethyl), 1-(2-acetoxyethyl), 1-(2-cyanoethyl) substituted 4-nitroimidazoles and 1-(2-hydroxyethyl)-2-nitroimidazole were unsuccessful. Treatment of 1-(2-acetoxyethyl)-5-nitro-2-imidazolecar-baldehyde(10) with hydroxylamine-O-sulfonic acid afforded a mixture of corresponding 2-carbonitrile (12) and 2-(N-hydroxy)carboximidamide (13). Hydrolysis of 10 with ethanolic hydrochloric acid yielded 8-ethoxy-5,6-dihydro-3-nitro-8H-imidazo[2,1-c] [1,4]oxazine (11) which, on subsequent reaction with hydroxylamine-O-sulfonic acid, afforded 1-(2-hydroxyethyl)-5-nitroimidazole-2-(N-hydroxy)carboximidamide (15). Reaction of 4(5)-nitroimidazole with chloropropionitrile produced a mixture of the isomeric 1-(2-cyanoethyl) substituted 4- and 5-nitroimidazoles. Treatment of 2,4(5)-dinitroímidazole with chloropropionitrile afforded a mixture of 4(5)-chloro-5(4)-nitroimidazole and 1-(2-cyanoethyl)-4-nitro-5-chloroimidazoIe. Reaction of nitroimidazoles with acrylonitrile in the presence of Triton B yielded the corresponding 1-(2-cyanoethyl) substituted derivatives.  相似文献   

10.
The reactions of carbethoxycarbene (:CH2-CO2Et, 2) with several acyclic enaminones (RCOCH=CR1NHR2, 3) lead to the unexpected formation of 2-Me, 3-CO2Et, 4-H, 5-R1-pyrroles 4 . Structural variations of the enaminones show that the structural fragments C(3)-CO2Et and C(2)-Me are provided by 2 and that the fragment C(5)-R1NHR2 originates from the enaminones 3 , while the RCO group from 3 is eliminated during the course of reaction. Reactions with cyclic and nitrogen-hindered enaminones do not lead to pyrrole formation but occur by simple insertion of 2 to the Cα-H bond.  相似文献   

11.
The regio‐ and stereoselectivity of cycloadditions of the nitrone 1a and the chiral, sugar‐derived nitrones 13a and 13b with 3‐(prop‐2‐enoyl)‐1,3‐oxazolidin‐2‐one ( 2 ) depends on the nature of the Lewis acid catalyst used. Addition of Lewis acid reverses the regioselectivity of the cycloaddition, and improves the anti‐diastereoselectivity in the case of chiral nitrones. The sterically favored isoxazolidin‐5‐yl‐substituted adducts 3, 4 , and 14 – 17 are produced as the major products in the absence of Lewis acid, while the electronically favored regioisomers with isoxazolidin‐4‐yl substituents ( 5, 6 , and 18 – 21 , respectively) are obtained as major products in the [Ti(OiPr)2Cl2] catalyzed reactions. The reactions of nitrone 13b with 2 in the presence of other Lewis acids such as ZnCl2, ZnBr2, ZnI2 and MgI2/I2 gave both regioisomeric pairs of the diastereoisomers, favoring the 4‐substituted congeners. The diastereoisomeric isoxazolidines 3a – 6a were reduced with NaBH4 in THF/H2O with subsequent desilylation to yield the separable diols 9 – 12 . Reduction of the diastereoisomeric isoxazolidines 19a and 18a afforded the chiral alcohols 23 and 22 , the latter of which was analyzed by X‐ray crystallography.  相似文献   

12.
New complexes of cadmium iodide with 1,3-bis[2-(diphenylphosphoryl)phenoxy]propane [CdL0I2], 1,2-bis[2-(diphenylphosphorylethyl)phenoxy]ethane [CdL1, 2I2], and 1,8-bis[2-(diphenylphosphorylethyl) phenoxy]-3,6-dioxaoctane [CdL3, 2I2] are synthesized and their IR spectra and crystal structures are studied. Electroanalytical characteristics of membranes of ion-selective electrodes based on L0, L1, 2, L3, 2, and known crown ethers are compared for cations of alkali, alkaline-earth, and transition metals. Ligand L3, 2 is the first podand with terminal diphenylphosphoryl fragments to exhibit selectivity with respect to the cadmium cation.  相似文献   

13.
Gas-phase reactions of dichlorogermylenes formed in the GeCl4-Si2Cl6 system with 2-chlorothiophene, 2,5-dichlorothiophene, 2,5-dichloro-3-methylthiophene, and 5-chloro-2-trimethylsilylthiophene is studied. It is shown that :GeCl2 is inserted into the chlorothiophene C-Cl bond to form the derivatives containing one or two trichlorogermyl groups. In the reaction with 5-chloro-2-trimethylsilylthiophene, the insertion of :GeCl2 into the C-Cl bond yielding 5-trichlorogermyl-2-trimethylsilylthiophene is accompanied by the exchange of the Me3Si group with the trichlorogermyl group to form 2,5-bis(trichlorogermyl)thiophene. The reactions occurring in the course of the synthesis of thienylchlorogermanes are considered.  相似文献   

14.
Summary.  Reactions of cyanomethanesulfonamides with aromatic aldehydes in the presence of AcOH and piperidine produced the addition products, the 1-cyano-2-arylethenesulfonamides, whereas reactions with benzonitrile yielded the 2-amino-1-cyano-2-phenylethenesulfonamides only when done in THF with BuLi. No addition products were isolated from the analogue reactions with 2-hydroxybenzaldehyde (salicylaldehyde). Instead, we obtained 2-imino-2H-chromene-3-sulfonamides with good to excellent yields. These 2H-chromene derivatives allowed a number of transformations, from which the reactions with orthoformates opened an approach to the hitherto unknown benzopyrano[3,2-e] [1,2,4]thiadiazine ring system.  相似文献   

15.
The electrochemical properties of single crystals of cerium fluoride alloyed with bivalent cations Sr2+, Ca2+, Ba2+, Sr2+ + Ca2+, Sr2+ + Ba2+, Ba2+ + Ca2+ and also with La3+ and La3+ + Ba2+ cations are studied using the dynamic voltammetry and impedance spectroscopy. The conductivity of symmetrical cells with Ag electrodes is determined using the method of impedance spectroscopy in the frequency range from 450 to 5 kHz at the temperatures from 20 to 100°C: for CeF3: Sr2+ (0.5 mol %) + Ba2+ (0.5 mol %), σ = σ0 exp[(?0.284 ± 0.005/kT]; for CeF3:Ca2+ (0.5 mol %) + Sr2+ (0.5 mol %), σ = σ0 exp[(?0.292 ± 0.017/kT]. The steady-state and dynamic voltammogams of symmetrical electrochemical cells with nonpolarizable reference electrodes and CeF3 single crystals alloyed with Sr2+, Ca2+, and Ba2+ bivalent cations exhibited ohmic polarization. For cells with CeF3 containing La3+ as an admixture, a hysteresis was observed, which could not be eliminated by chemical and electrochemical treatment of crystals. In the dynamic voltammetric curves of asymmetric cells with nonpolarizable and silver electrode and CeF3 crystals alloyed with Sr2+, Ca2+, and Ba2+, a range of ideal polarizability (from 0 to ~?2.7 V), and also cerium redox processes and silver fluorination-boundary regeneration were observed. In the dynamic voltammetric curves of asymmetric cells with CeF3 containing La3+ admixture, no range of ideal polarizability was observed; however, the reactions of silver fluorination and reduction of solid-electrolyte cerium were well pronounced at the corresponding potentials.  相似文献   

16.
Reactions of diesel soot and NOx with and without O2 were carried out over CuFe2O4 catalyst. The ignition temperature of soot with the NO+O2 feed was lower than that in O2 or NO but close to those in NO2 and NO2+O2, indicating the implication of NO2 especially in decreasing the ignition temperature. On the other hand, the reduction of NOx into N2 was enhanced by coexisting O2. Based on these results and mechanisms of O2-soot and NO-soot reactions, the possible reaction mechanism of the simultaneous NOx-soot removal with the NO+O2 feed has been proposed.  相似文献   

17.
Chloro-dimethylamino-phenyl-p-tolylthio-phosphonium chloride 2 , dimethylamino-diphenyl-p-tolylthio-phosphonium chloride 3 , bis(diethylamino)-dimethylamino-p-tolylthiophosphonium chloride 4 and tert-butyl-dimethylamino-phenyl-p-tolylthio-phosphonium chloride 5 were prepared by the reaction of N,N-dimethylamino-p-tolylsulfenamide 1 with PhPCl2, Ph2PCl, (Et2N)2PCl and tBu(Ph)PCl, respectively. The reaction of N,N′-dimethyl-N,N′-bis(trimethylsilyl)urea 9 and N-methyl-N′-phenyl-N,N′-bis(trimethylsilyl)urea 10 with phenylsulfenyl chloride 6 or p-nitrophenylsulfenyl chloride 8 furnished the N-arylthio-N,N′-diorgano-N′-(trimethylsilyl)-ureas 11 – 14 . The reaction of 11 – 14 and of the previously known compounds 15 and 16 with MePCl2, ClCH2PCl2, tBuPCl2 and PhPCl2 resulted in the formation of the 2-arylthio-2-chloro-1,2,3-triorgano-1,3,2λ5-diazaphosphetidin-4-ones 17 – 26 . 1,3-Dimethyl-2-(1,1,1,3,3,3-hexafluoro-2-propoxy)-2-phenyl-2-phenylthio-1,3,2λ5-diazaphosphetidin-4-one 29 and the 2-arylthio-1,3-dimethyl-2-(p-nitrophenoxy)-2-organyl-1,3,2λ5-diazaphosphetidin-4-ones 30 – 32 were obtained in the reactions of compounds 17, 24 and 27 with 1,1,1,3,3,3-hexafluoro-2-propanol or p-nitrophenol in the presence of triethylamine. The reaction of compound 21 with thiophenol in the presence of triethylamine resulted in a mixture of products, from which 1,3,4,5,7-pentamethyl-1,3,5,7-tetraaza-4λ5-phosphaspiro[3.3]heptan-2,6-dione 33 was isolated. The identity and structure of all the new compounds were established by 1H-, 13C- and 31P-NMR spectroscopy and by elemental analysis. A possible mechanism of reaction of sulfenamides with compounds of trivalent phosphorus is discussed. For the compounds 5a, 32 and 33 X-ray structure analyses were conducted. The cation of compound 5a involves four-coordinate phosphorus (essentially tetrahedral geometry) and is a rare example of a P–S single bond in such a system (P–S 207.37(9) pm). In 32 the geometry at phosphorus is distorted trigonal bipyramidal, with axial positions occupied by oxygen and nitrogen atoms. In the spirophosphorane 33 the geometry at phosphorus is intermediate between trigonal bipyramidal and square pyramidal, with essentially planar four-membered rings.  相似文献   

18.
By using frontier‐molecular‐orbital and electrostatic (nucleophilic) interactions as well as relaxed potential‐energy surface scans, it is shown that the initial step in the oxygen‐atom transfer (OAT) reaction of [MoVIO2‐(S2C2Me2)SMe]?1 ( 1 ) and [MoVIO2‐{(S2C2(CN)2}2]2? ( 2 ) with HSO3? takes place by oxoanionic binding of the substrate to the MoVI center with the formation of a stable Michaelis complex. The gas‐phase and solvent‐corrected enthalpy profile with fully optimized minima and transition states for the OAT reaction of 1 and 2 with HSO3? showed the release of reaction energy for both complexes. The optimized geometries of 1 and 2 in the respective enzyme–substrate complexes showed a common feature with the participation of hydrogen bonding of the substrate with the axial (spectator) oxo group in the subsequent formation of the six‐membered MoO2HOS transition state. The enzyme–substrate complex of 2 shows heptacoordination as proposed earlier, although the trans (to axial oxo)‐Mo? S(dithiolene) bond is elongated to 2.948 Å.  相似文献   

19.
The kinetics and mechanism of Cl-atom-initiated reactions of CHO? CHO were studied using the FTIR detection method to monitor the photolysis of Cl2–CHO? CHO mixtures in 700 torr of N2–O2 diluent at 298 ± 2 K. The observed product distribution in the [O2] pressure of 0–700 torr combined with relative rate measurements provide evidence that: (1) the primary step is Cl + CHO? CHO → HCl + CHO? CO with a rate constant of [3.8 ± 0.3(σ)] × 10?11 cm3 molecule?1 s?1; (2) the primary product CHO? CO unimolecularly dissociates to CHO and CO with an estimated lifetime of ≤ca. 1 × 10?7 s; (3) alternatively, the CHO? CO reacts with O2 leading to the formation of CO, CO2, and most likely the HO radical, but no stable products containing two carbon atoms; (4) the HO2 radical, formed in the secondary reaction CHO + O2 → HO2 + CO, reacts with the CHO? CHO with a rate constant ca. 5 × 10?16 cm3 molecule?1 s?1 to form HCOOH and a new transient product resembling that detected previously in the HO2 reaction with HCHO.  相似文献   

20.
The enthalpy of the reduction of UO2F2 with hydrogen was obtained from quantitative DTA measurements with a linear heating rate and under isothermal conditions, and the thermodynamic data on UO2F, formed as a stable intermediate in the reduction of UO2F2 to UO2, are also presented. The advantages of isothermal DTA in the reduction of U3O8 to UO2 could be demonstrated.  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