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 photocatalytic transformation of dexamethasone and the formation of its intermediate compounds have been studied using titanium dioxide as a photocatalyst. The degradation of dexamethasone occurs easily through the formation of several hydroxy derivatives whose characterization has been made by HPLC/MS/MS. Even if both oxidative and reductive processes can be operating, only oxidative products have been identified in air saturated aqueous suspensions. A pattern of reaction pathways accounting for the observed intermediates is proposed. The obtained experimental evidence may be rationalized postulating the existence of a double initial mechanism. A single oxidation step resulting from the attack by one ·OH radical leading to the formation of five hydroxy-derivatives and a concomitant attack involving two ·OH radicals leading to the hydroxylation of the quinoid moiety of the molecule.  相似文献   

2.
Dimethyl carbonate (DMC) synthesis reaction by oxidative carbonylation of methanol has been studied using vapor phase flow reaction system in the presence of Cu-based catalysts. A series of Cu-based catalysts were prepared by the conventional impregnation method using activated carbon (AC) as support. The effect of various promoters and reaction conditions on the catalytic reactivities were intensively evaluated in terms of methanol conversion and DMC selectivity. The morphological analysis by X-ray diffraction and SEM was also conducted in order to characterize the emloyed catalysts. Regardless of catalyst compositions, the optimal reaction temperature for oxidative carbonylation of methanol was found to be around 120–130°C. The reaction rate was too slow below 100°C, while too much by-products was produced above 150°C. Among the various catalysts employed, CuCl2/NaOH/AC catalyst with the molar ratio of OH/Cu=0.5–1.0, has shown the best catalytic performance, which appears to have a strong relationship with the formation of intermediate species, Cu2(OH)3Cl.  相似文献   

3.
Y. Ogata  A. Kawasaki  K. Nakagawa 《Tetrahedron》1964,20(12):2755-2761
The rates of the formation of benzoguanamine from dicyandiamide and benzonitrile in the presence of a basic catalyst, have been measured by means of UV spectrophotometry. The rate is expressed as ν = k × [dicyandiamide]·[benzonitrile]·[KOH]. Since the addition of a small amount of cyanamide retards the rate, cyanamide cannot be an intermediate. Application of the rate data observed with p-methyl-, methoxy-, and chlorobenzonitriles to the Hammett's equation gives a positive value of +1·82. A probable mechanism involving a rate-determining attack of the conjugate base of dicyandiamide on benzonitrile is suggested and discussed.  相似文献   

4.
The oxidation of white oil solutions of dibenzothiophene (DBT) by aqueous hydrogen peroxide-acetic acid solutions was studied kinetically at 50–100°. Under these conditions, the rate of DBT oxidation was found to be first order in acetic acid, second order in hydrogen peroxide, and inversely proportional to the water concentration. The activation energy between 50–100° in 64·5% acetic acid was 14 kcal/mole. We have also found that the monoxide is oxidized about 1·4 times faster than DBT. A mechanism consistent with the kinetic data has been postulated. The rate-determining step appears to be attack of a peracetic acid-hydrogen peroxide dimer on the sulfur atom of DBT.  相似文献   

5.
Cukrowska E  Cukrowski I 《Talanta》1998,47(5):2799-1189
The ligand monoaza-12-crown-4 ether (A12C4) was studied in aqueous solution at 298 K and an ionic strength of 0.5 mol dm−3 in the presence of an excess of sodium ion (0.5 mol dm−3 NaNO3). The protonation constant of A12C4, determined by glass electrode potentiometry (GEP) in the same background electrolyte, was found to be log K=9.36±0.03. Polarographic experimental and calculated complex formation curves (ECFC and CCFC) for labile metal–ligand systems, studied at a fixed total ligand (LT) to total metal (MT) concentration ratio and varied pH, were used for the modelling of the metal species formed and the refinement of their stability constants. The metal–ligand model and formation constants are optimised by solving mass-balance equations written for the assumed model and by fitting the CCFC to the ECFC. The CCFC can be generated for any metal–ligand model, including polynuclear metal species, for any LT:MT ratio, and for more than one ligand competing in the complex formation reaction. Three lead complexes with the ligand A12C4, viz. PbL2+, PbL(OH)+ and PbL(OH)2, were found and their overall stability constants from differential pulse polarography (DPP), as log β, were estimated to be 3.75±0.03, 9.30±0.05 and 12.70±0.05, respectively. Two copper complexes CuL2+ and CuL(OH)2 are reported and their stability constants (from DPP) were estimated to be 6.00±0.05 and 21.77±0.1, respectively. Two cadmium complexes CdL2+ and CdL(OH)+ are reported. The stability constant for CdL2+ was estimated from DPP and GEP as 2.80±0.05 and 2.68±0.03 (the latter value was obtained from a few potentiometric experimental points), respectively, and the stability constant for CdL(OH)+ from DPP was estimated to be 7.88±0.05. GEP could not be used for the stability constants determination of other metal complexes studied because of precipitation occurring prior the completion of a complex formation reaction.  相似文献   

6.
Copper(II) reacts with rac-5,5,7,12,12,14-hexamethyl-l,4,8,11-tetraazacyclotetradecane (tet-b) in strongly basic aqueous media to give [Cu(tet-b) (OH) (blue)]+ which contains trigonal bipyramidally co-ordinated Cu2+ with the tet-b ligand in its most stable, folded form. The kinetics of formation of this blue complex have been studied at 25.0° ± 0.1°C using the stopped-flow technique. Second-bond formation is proposed as the rate-determining step for tet-b reaction with Cu(OH)-3 and Cu(OH)2-4. Possible mechanisms for the reaction and the steric effects resulting from the methyl groups on the alkyl backbone of the macrocyclic ligand are considered.  相似文献   

7.
Hydrated strontium borate, SrB4O7·3H2O, has been synthesized and characterized by XRD, FT-IR, DTA-TG and chemical analysis. The molar enthalpy of solution of SrB4O7·3H2O in 1 mol dm−3 HCl(aq) was measured to be (21.15 ± 0.29) kJ mol−1. With incorporation of the previously determined enthalpies of solution of Sr(OH)2·8H2O(s) in [HCl(aq) + H3BO3(aq)] and H3BO3 in HCl(aq), and the enthalpies of formation of H2O(l), Sr(OH)2·8H2O(s) and H3BO3(s), the enthalpy of formation of SrB4O7·3H2O was found to be −(4286.7 ± 3.3) kJ mol−1.  相似文献   

8.
Rate constants have been measured at 296 ± 2 K for the gas‐phase reactions of camphor with OH radicals, NO3 radicals, and O3. Using relative rate methods, the rate constants for the OH radical and NO3 radical reactions were (4.6 ± 1.2) × 10−12 cm3 molecule−1 s−1 and <3 × 10−16 cm3 molecule−1 s−1, respectively, where the indicated error in the OH radical reaction rate constant includes the estimated overall uncertainty in the rate constant for the reference compound. An upper limit to the rate constant for the O3 reaction of <7 × 10−20 cm3 molecule−1 s−1 was also determined. The dominant tropospheric loss process for camphor is calculated to be by reaction with the OH radical. Acetone was identified and quantified as a product of the OH radical reaction by gas chromatography, with a formation yield of 0.29 ± 0.04. In situ atmospheric pressure ionization tandem mass spectrometry (API‐MS) analyses indicated the formation of additional products of molecular weight 166 (dicarbonyl), 182 (hydroxydicarbonyl), 186, 187, 213 (carbonyl‐nitrate), 229 (hydroxycarbonyl‐nitrate), and 243. A reaction mechanism leading to the formation of acetone is presented, as are pathways for the formation of several of the additional products observed by API‐MS. © 2000 John Wiley and Sons, Inc. Int J Chem Kinet 33: 56–63, 2001  相似文献   

9.
Homogeneous Fenton reaction (H2O2/Fe2+ system) using significantly substoichiometric concentrations of H2O2 oxidant to oxidize phenol was characterized focusing on the formation of stable aromatic intermediates. Beyond the most abundant benzenediols, the pattern of aromatic intermediates was chiefly characterized by hydroxylated biphenyls and diphenyl ethers with different degrees of hydroxylation. Hydroxylated dibenzofurans (DBF), p,p′-dioxins, as well as highly condensed aromatic intermediates including polyols of polycyclic aromatic hydrocarbons (PAHs), could also be detected, but in lower concentrations. The formation of aromatic intermediates could be predicted on the basis of oxidative coupling reactions of resonance-stabilized radicals generated by the attack of the highly reactive hydroxyl radicals (OH*) on phenol. GC/MS identification of oxidative coupling intermediates was performed after derivatization of the solvent extracts. Derivatization reactions included silylation to give TMS (trimethylsilyl) ethers, as well as single-step extractive acetylation using acetic anhydride in alkaline aqueous solutions (pH 10.5) to give acetates. Solvent extraction of aqueous solutions, a prerequisite to generate TMS ethers, caused strong discrimination of polyols due to their low distribution coefficients in non-polar solvents. This discrimination could be overcome by extracting the in-situ formed acetates of the intermediates. Extractive acetylation allowed the detection of tri-, tetra-, and penta-hydroxylated aromatic intermediates generated by Fenton oxidation processes, which have been overlooked upto now. Thus, extractive acetylation to detect stable aromatic intermediates covering a wide range of hydroxylation degrees can foster the understanding, monitoring, and management of advanced oxidation processes, especially in the field of wastewater treatment.  相似文献   

10.
We have chosen 3-ethoxypropan-1-ol (EtOPrOH) as a typical short-carbon-chain ether–alcohol used as industrial solvent and have analyzed the degradation products resulting from its attack by OH radicals generated by the photo-Fenton reaction. The laboratory conditions were representative of those found in tropospheric water droplets. Twelve products have been identified by use of GC–MS analysis, either directly or after extraction by SPME fibers, and by HPLC–UV analysis with a special column for carboxylic acids and after reaction of carbonyl groups with 2,4-dinitrophenylhydrazine. These products contain one to three carbon atoms (instead of five in EtOPrOH), among which one or two are oxidized. According to the reaction pathways proposed, seven products—including methanal—can result from attack by one OH only, three products imply attack by a second OH, as expected from their higher oxidation number, and it is suggested that reaction between two organic radicals is needed for formation of only two products. The relevance of this investigation to the fate of EtOPrOH and similar ether–alcohols in the troposphere is briefly discussed.  相似文献   

11.
Phase diagrams for the system nitrocellulose/tetrahydrofuran/ethanol have been mapped out at room temperature, for two different degrees of substitution of the polymer (2·32 and 2·72). The phases formed by particular compositions were identified by using polarized light microscopy. There is a range of compositions which separate into an isotropic phase and a liquid-crystalline phase; the concentrations involved are qualitatively in agreement with Flory's theory of phase separation in a solution of rigid rod-like molecules. Our plotted room temperature sections of the phase diagrams obey the Gibbs phase rule at all points. We report evidence for the formation of a crystallosolvate of nitrocellulose (degree of substitution 2·72) and ethanol. The crystallosolvate formation only appears to be kinetically possible if there is initially also sufficient tetrahydrofuran present to dissolve the polymer.  相似文献   

12.
The rates of the reactions of hydroxyl radicals (OH) with styrene, α-methylstyrene, and β-methylstyrene have been measured by irradiating mixtures of these aromatic olefins and NO in an environmental chamber at 298 K. Experimental conditions were used whereby the competition of ozone with OH in oxidizing the hydrocarbons could be considered negligible. The rate constant values, obtained by a relative method using isooctane as reference hydrocarbon, are: styrene (5.3 ± 0.5) × 10?11 cm3/molec·s, α-methylstyrene (5.3 ± 0.6) × 10?11 cm3/molec·s, and β-methylstyrene (6.0 ± 0.6) × 10?11 cm3/molec·s. A simplified kinetic treatment of the experimental data shows that styrene and β-methylstyrene are stoichiometrically converted to benzaldehyde, suggesting that OH attack occurs only on the aliphatic moiety of the aromatic olefins. Benzaldehyde was observed to undergo consecutive oxidation by OH, and its maximum formation yield was about 60%. A reaction mechanism is proposed where the primary rate-determining OH attack leads to the formation of 1-hydroxy-2-phenyl-2-ethenyl radicals, from which benzaldehyde is formed through fast intermediate reactions.  相似文献   

13.
A number of isomeric N-benzylbenzalimine palladium(II) complexes of the type [P ·CH2Ph]2 (with C=N endo to the palladocycle) and [P =C(CH3Ph]2 (with C=N exo to the palladocycle), have been prepared and charcterised by 1H and 13C NMR methods. The crystal structures of two analogous monomeric acac complexes, synthesized independently by oxidative addition of o-BrC6H4CH2N=CH · Ph to Ph to Pd(dibenzylideneacetone)2 have also been determined. These are [P · CH2Ph)] (15a) and [P =CHPh)] (20a). Crystals of 15a are monoclinic, space group P21/a with Z = 4 in a cell of dimensions a 10.286(2), b 11.902(3), c 13.895(5) Å, β 93.52(2)° while 20a is monoclinic, space group P21/c with Z = 8 and a 10.353(3), b 20.600(5), c 16.545(7) Å, β 92.14(3)°. The structures 15a and 20a were refined to residuals R = 0.041 and 0.055 for 1661 and 2525 observed reflections respectively.  相似文献   

14.
Huitink GM  Poe DP  Diehl H 《Talanta》1974,21(12):1221-1229
A satisfactory method for the preparation of Calcein Blue has been devised. Elemental analysis, equivalent weight by neutralization, and the NMR spectrum show the compound to be 4-methylumbelliferone-8-methyleneiminodiacetic acid·0·25H2O. The ultraviolet absorbance and fluorescence have been studied as a function of pH and, combined with potentiometric titration and solubility date, have yielded for the acid dissociation constants the values pK1 = 3·0, pK2 = 6·9, and pK3 = 11·3. These acid functions are identified respectively as carboxyl, phenol, and ammonium ion, the free Calcein Blue being a zwitter-ion. Calcein Blue fluoresces in both acidic and basic solution when excited at a suitable wavelength. The fluorescence of the doubly-charged anion formed on the neutralization of the phenol group, when excited at 360 nm, reaches a maximum at pH 9, and decreases to zero with the neutralization of the ammonium ion; the wavelength of maximum emission is 455 nm. In the presence of calcium, the fluorescence increases with alkalinity up to pH 9 and then remains constant. The calcium derivative is a 1:1 compound, formation constant 107·1. The fluorescence of Calcein Blue at all pH values is quenched by copper(II). The calcium derivative is changed on standing in highly alkaline solution, presumably by ring opening, to another fluorescent material; thus Calcein Blue, although satisfactory as an indicator, is not useful for the direct fluorometric determination of calcium.  相似文献   

15.
The 1,3-dithia-2-boryl[3]ferrocenophanes (I-VIII) are obtained from the reaction of ferrocene-1,1′-dithiol (A) with dihalogenorganylboranes by elimination of the hydrogenhalides. I-VIII can be also prepared by starting from Fe(C5H4SLi)2 · 2THF (B) inplace of A, but in lower yields. The air-sensitive [3] ferrocenophanes have been characterized analytically and spectroscopically (MS; NMR: 1H, 11B, 13C).  相似文献   

16.
G. Favero  U. Russo  M. Vidali  B. Zarli 《Polyhedron》1988,7(24):2703-2707
With the binucleating ligand 1-oxy-2,6-di [(N,N-biscarboxymethyl)aminomethyl]- 4-chlorobenzol (H5L) complexes of formulae FeH2L · 2H2O; FeH3L(C1O4) · H2O; Fe2L(OH) · 2H2O; M2HL · nH2O (M = Co, Cu, n = 2; M = Ni, n = 4); FeCuL · 3H2O; FeCrL(OH) · 3H2O were prepared and characterized by elemental analysis, IR and electronic spectra and magnetic moment determinations. In addition, thermal analysis data of the complexes and Mössbauer effect spectra of the iron containing complexes are also given and discussed.  相似文献   

17.
《Tetrahedron》1961,16(1-4):139-145
The OH proton resonance spectra of tertiary alcohols related to 3-buten-1-ol and phenethyl alcohol in carbon tetrachloride have been measured in wide range of concentration at 56·4 Mc/sec. The resonance position rises steadily to the higher applied magnetic field on dilution, owing to the dissociation of the associated alcohols. The chemical shifts at infinite dilution (δ0), which are obtained by extrapolating the δ versus concentration curves and concerned with the pure monomer, are characteristic for the respective alcohols. The results are discussed with reference to the data from the infra-red spectroscopy and the δi of the hydroxyl group interacting with the ethylenic and the benzenoid π-electrons are obtained as 3·19 and 3·50 parts per million, respectively, from the proton of water used as an external standard. The difference may be attributed either to the higher basicity of the ethylenic compounds or to the diamagnetic effect by the ring current of the phenyl group.  相似文献   

18.
Reactivity ratios for the radical copolymerisation of the methylmethacrylate (MMA) (1)/methyl isopropenyl ketone (MIK) (2) system have been evaluated at 60°C as r1=0·97; r2=1·09. Copolymers with MIK contents from 0 to 15% have been prepared. Films were exposed to sunlight under environmental conditions at ground level and buried under-ground at a depth of 9 cm and their photo-degradation, as measured by chain scissions and mechanical properties, was followed with exposure time. For the specimens at ground level, a clear dependence of degradation on sunlight exposure time and MIK content is observed, similar to that observed under laboratory conditions, whether in film or in solution. Negligible degradation was observed over a 2-year period in the buried specimens. Laboratory biodegradation tests seem to indicate that attack by microorganisms starts at a polymer molecular weight of about 20 000.  相似文献   

19.
Y. Ogata  Y. Kosugi  K. Tomizawa 《Tetrahedron》1970,26(24):5939-5944
Vitamin A,dissolved in liquid paraffin, is stable below room temperature, but suffers oxidative decomposition at 80°, giving its epoxide as the main product. The rate of decomposition of vitamin A (VA) at 80° in the presence of oxygen (partial pressure of p) and a small amount of cobaltous stearate (Co) is expressed as: d[VA]/dt = 3·15 × 10−1:[VA][Co] p1·48 + 1·51 × 10−5[VA]p0·33 + 0·33 × 10−5[VA], where the last term represents the spontaneous thermal decomposition.  相似文献   

20.
Dibenzothiophene (1) suffers surface-catalyzed oxidation under CI(O2) conditions. While in the positive mode M is the only major ion and those of the oxidation products are of minor importance in the negative mode, essentially only ions stemming from oxidized species can be seen, the sulfone ion [M + O2]−· being the most important one. The ion m/z 184 previously attributed to M−· is actually the anion of 2-sulfobenzoic acid cyclic anhydride. Structures for the various oxidation products are proposed and the mechanisms leading to their formation are discussed.  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