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 781 毫秒
1.
The gas-phase partial oxidation of benzene and toluene to phenol and to cresols over Fe-containing ZSM-5 zeolites prepared by various methods was studied using N2O as oxidant. Over FeZSM-5 synthesized by either isomorphous substitution or ion exchange the phenol selectivity reached nearly 100% at 573 K, however, only 25-30% cresol (mainly p-cresol) selectivity was observed in the oxidation of toluene because of the formation of benzaldehyde and benzoic acid in the fast oxidation of the methyl group.  相似文献   

2.
An experimental and modeling study of irradiated toluene–NOx–air, toluene–benzaldehyde–NOx–air, and cresol–NOx–air mixtures at part-per-million concentrations has been carried out. These mixtures were irradiated at 303 ± 1 K in a 5800-liter Teflon-lined, evacuable environmental chamber, with temperature, humidity, light intensity, spectral distribution, and the concentrations of O3, NO, NO2, toluene, PAN, formaldehyde, benzaldehyde, o-cresol, m-nitrotoluene, and methyl nitrate beingmonitored as a function of time. For the toluene and toluene–benzaldehyde–NOx–air runs a variety of initial reactant concentrations were investigated. Cresol–NOx–air runs were observed to be much less reactive in terms of O3 formation and NO to NO2 conversion rates than toluene–NOx–air runs, with the relative reactivity of the cresol isomers being in the order meta » ortho > para. The addition of benzaldehyde to toluene–NOx–air mixtures decreased the reactivity, in agreement with previous studies. Alternative mechanistic pathways for the NOx photooxidations of aromaticsystems in general are discussed, and the effects of varying these mechanistic alternatives on the model predictions for the toluene and o-cresol–NOx–air systems are examined. Fits of the calculations to most of the experimental concentration–time profiles could be obtained to within the experimental uncertainty for two of the mechanistic options considered. In both cases it is assumed that (1) O2 adds to the OH–toluene adduct ~75% of the time forming, after a further addition of O2, a C7 bicyclic peroxy radical, and (2) this C7 bicyclic peroxy radical reacts with NO ~75% of the time to ultimately form α-dicarbonyls and conjugated γ-dicarbonyls (e.g., methylglyoxal + 2-butene-1,4-dial) and ~25% of the time to form organic nitrates. The major uncertainties in the mechanisms concern (1) the structure of the bicyclicperoxy intermediate, and (2) the γ-dicarbonyl photooxidation mechanism. Good fits to the o-cresol concentration–time profiles in the toluene–NOx runs are obtained if it is assumed that o7-cresol reacts rapidly with NO3 radicals. However, it is observed that the model underpredicts nitrotoluene yields by a factor of ~10, but this is in any case a minor product. It is concluded that further experimental work will be required toadequately validate the assumptions incorporated in the aromatic photooxidation mechanisms presented here.  相似文献   

3.
Acute toxicity of cresols to both Pseudomonas I and II was estimated by an initial oxygen uptake method. Inhibition studies of toluene and cresols on the oxidation of either benzoate by Pseudomnas I or phenol by Pseudomonas II were analyzed and expressed as oxygen uptake rates. Double reciprocal plots for the inhibiton by cresols of oxygen uptake in Pseudomonas, two physical constants, Vmaxi and Ki, were obtained. The Vmaxi of o?, m? and p-cresol were 80%, 81% and 57% of Vmax in Pseudomnas I, and 10%, 25% and 36% in Pseudomonas II, respectively. Thus, the toxicity to Pseudomonas I decreases in the order p- > o- ≥ m-cresol, whereas to Pseudomonas II, the order is changed to o- > m- > p-cresol. This difference in the toxicity order is probably due to the allosteric effect of p-cresol towards Pseudomonas II. Inasmuch as most compounds inhibit noncompetively, the relative toxicity of different compounds can be estimated by a new toxicity parameter RI (relative inhibition) which is defined as 100/Ki. By comparing the RI value of each compound, the toxicity to Pseudomonas I decreases in the order m-chlorophenol > p-cresol > p-chlorophenol > o-cresol ≥ m-cresol > o-chlorophenol > toluene > phenol.  相似文献   

4.
BF3·OEt2-catalyzed methyl group migration of 4-methyl-4-tert-butyldioxycyclohexadienone, which is obtained by ruthenium-catalyzed oxidation of p-cresol with tert-butyl hydroperoxide, in hexafluoro-2-propanol/toluene gave toluquinone efficiently. The reaction can be applied to the regio-selective short-step syntheses of vitamin K3 and vitamin K1 from p-cresol.  相似文献   

5.
《Fluid Phase Equilibria》1999,157(1):53-79
Phase equilibria in binary and ternary systems containing o-cresol, p-cresol, carbon dioxide, and ethanol have been investigated experimentally at temperatures between 323.15 K and 473.15 K and pressures ranging from 10 MPa to 35 MPa. The experimental results provide a systematic basis of phase equilibrium data, yielding the effect of temperature on the influence of the position of the methyl groups of cresols that are in phase equilibria with carbon dioxide. Based on the different solubilities of the cresol isomers in carbon dioxide, the separation of o-cresol and p-cresol was investigated. The dependence of the separation factor between both cresol isomers on concentration, temperature, and pressure is obtained from experiments in the ternary system, o-cresol+p-cresol+carbon dioxide. The influence of ethanol added to each of the binary systems, cresol isomer+carbon dioxide, in order to enhance the solubility of the cresols in the carbon dioxide-rich phase is also shown. The experimental data have been correlated using seven different equations of state, whereof four explicitly account for intermolecular association: Statistical Association Fluid Theory (SAFT) by Chapman, Gubbins, Huang and Radosz, the SAFT modification by Pfohl and Brunner for near-critical fluids, a modified cubic-plus-association equation of state (CPA EOS) according to the ideas by Tassios et al., and one of the EOS by Anderko. The mixing rule proposed by Mathias, Klotz, and Prausnitz, with two binary interaction parameters per binary system influencing intermolecular attractive forces, is used for all EOS as a basis for an objective comparison of the EOS.  相似文献   

6.
The solvent effect on the kinetic energy distribution of p-cresol molecule was investigated by quantum-chemical (QM) method and molecular dynamics (MD) simulations, then the consequences were checked experimentally by photoluminescence (PL) and differential scanning calorimetry (DSC) methods. Results of QM calculations highlight the coupling of two vibrational normalmodes of p-cresol molecule in the presence of ethanol while no similar coupling was observed in methanol. MD simulations show that the normalmode coupling in ethanol is more pronounced at higher temperature and it is preferably based on the molecular friction of the cresol molecule with its environment. The theoretical observations were also proved experimentally. The dissociation rate of calixarene–phenol complexes were measured by DSC method. The decreased dissociation rate of the calixarene–phenol complexes observed in ethanol reflects the increased motion of the methyl groups of cresol units of calixarene in the ethanol solvent, a property which was predicted by the theoretical results. Our findings are applicable to many areas of chemistry where the formation and dissociation rates play important role: e.g., in the development of chemical molecular sensors or developing molecular containers for drugs towards pharmaceutical applications.  相似文献   

7.
Abstract

A simple, rapid and sensitive titrimetric method with amplification has been worked out for the determination of 50–2000μg of o-cresol, m-cresol or p-cresol. It is based on bromination of these compounds with bromine water to form the corresponding hypobromites, which liberate an equivalent amount of iodine when treated with iodide. The sensitivity of the method has been increased using the Leipert amplification procedure. The coefficient of variation does not exceed 1.5% for above 500μg of the cresol, but increases to 2.2% at the 50μg-level.  相似文献   

8.
A novel course of a phenylation reaction of 1,5-dihydroxy-4,8-diaminoanthraquin-one-2,6-disulfonic acid with cresols . The arylation of 1,5-dihydroxy-4,8-diaminoanthraquinone-2,6-disulfonic acid with m-cresol in conc. sulfuric acid gives in the presence of boric acid a mixture of monosulfonic acids which differ in the substitution of the m-cresol moiety. The main product ( 8 , 95%) is substituted at the p-position to the methyl group, the side product ( 12 , 5%) at the p-position to the OH group. The monosulfonic acid 8 , which could not be isolated is further sulfonated under the reaction conditions to the disulfonic acid 9 . In the case of o-cresol, the cresol moiety is substituted in the p-position ( 16 ) to OH group and in the case of p-cresol in the o-position ( 20 ) to OH group. The obtained monosulfonic acids 16 and 20 resp. are partially sulfonated further under the reaction conditions. The new structures are elucidated by 1H- and 13C-NMR. spectroscopy and the pattern of arylation reaction with phenol is discussed.  相似文献   

9.
Density, viscosity and surface tension of nine binary liquid systems: ethanoic acid, propanoic acid and butanoic acid with o-cresol, m-cresol and p-cresol have been determined at 298.15, 308.15 and 318.15 K over the complete compositional range. From the experimental results the excess values of molar volume (V E), viscosity (η E), Gibbs free energy for the activation of flow (ΔG E) and surface tension (σ E) were evaluated. The excess values were fitted to the Redlich–Kister type equation using a nonlinear regression technique. The Grunberg–Nissan parameter, d, was also calculated. From the sign and magnitude of the V E, η E, ΔG E, σ E, and d values, it is concluded that specific interactions are present in all of the nine binary mixtures under study. V E is negative for carboxylic acid–cresol mixtures at all temperatures and over the entire composition range. The values of η E, ΔG E and σ E are positive over the whole range of composition and increase with increasing temperature at a constant mole fraction of the carboxylic acid, confirming the existence of specific interactions in these binary mixtures. Further, the viscosity data of the binary systems were fitted to various theoretical/empirical models. The binary viscosity data is well represented by the Auslander model. Surface tension data were fitted to various theoretical/empirical models. The binary mixture surface tension data are well represented by the model given by Zihao and Jufu.  相似文献   

10.
An experimental investigation of the gas-phase photooxidation of toluene–NOx–air mixtures at part-per-million concentrations has been carried out in a 65-m3, outdoor smog chamber to assess our understanding of the atmospheric chemistry of toluene. In addition, six CO? NOx–air irradiations were conducted to characterize the chamber with regard to any wall radical sources. Measured parameters in the toluene–NOx experiments included O3, NO, NO2, HNO3, peroxyacetyl nitrate (PAN), CO, toluene, benzaldehyde, o-cresol, m-nitrotoluene, peroxybenzoyl nitrate (PBZN), temperature, relative humidity, aerosol size distributions, and particulate organic carbon. Predictions of the reaction mechanism of Leone and Seinfeld [7] are found to be in good agreement with the data under a variety of initial conditions. Additional simulations are used to investigate various mechanistic pathways in areas where our understanding of toluene chemistry is still incomplete.  相似文献   

11.
Preparation of aromatic alcohols and aldehydes by oxidation of p-cresol with ozone in acetic anhydride in the presence of sulfuric acid, manganese acetate, and potassium bromide was studied. The optimal oxidation conditions were determined.  相似文献   

12.
All stereoisomers of 2-allyl-3-hydroxy-2-methylcyclopentanones 2-5 were prepared in high conversion and in an optically pure form by microbial reduction and oxidation. The reduction of symmetric diketone 1 by Geotrichum candidum NBRC 4597 under anaerobic conditions gave 2 in 83% yield (98% conversion), >99% de, and >99% ee, whereas the reduction of 1 by G. candidum NBRC 5767 under aerobic conditions gave 3 in 75% yield (99% conversion), >99% de, and >99% ee. Oxidation of meso-diol 6 by G. candidum NBRC 5767 under aerobic conditions afforded 4 in 83% yield (99% conversion) and >99% ee, while oxidation of meso-diol 7 by Mucor heimalis IAM 6095 in the presence of cyclohexanone as a co-oxidant afforded 5 in 68% yield (75% conversion) and >99% ee.  相似文献   

13.
The kinetics of the liquid-phase catalytic oxidation of para-cresol with an ozone-air mixture in the presence of manganese(II) acetate is reported. In an acetic anhydride medium, para-cresol reacts with ozone as para-cresyl acetate, which is formed at the instant the solution to be oxidized is prepared. Under these conditions, the major oxidation products are para-acetoxybenzyl acetate (63.5%) and para-acetoxybenzylidene diacetate (13.7%). The effect of the managanese(II) acetate concentration on the traction selectivity with respect to the oxidation of the methyl group of the substrate is reported. A mechanism consistent with the experimental data available on this catalytic redox reaction is suggested.  相似文献   

14.
The γcmc values of CTAB-SDS decrease from 63.67 mN/m at 10‡C to 36.38 mN/m at 90‡C, slightly lower than those of either CTAB or SDS. Correspondingly, the CMC of CTAB-SDS decreases almost by half. The increase of surface activity of CTAB-SDS can be attributed to the relatively weak electrostatic interaction at high temperature, which is supported by the increase of solubility of CTAB-SDS with rise in temperature. Catalytic effect on oxidation of toluene derivatives with potassium permanganate follows the order CTAB-SDS > SDS > CTAB. This is not caused by the dissociative effect of CTAB-SDS with low surface activity at low temperature, as seen from the fact that almost all oxidative products can be retrieved for different toluene derivatives and surfactants by mimicking the conditions of reaction. In the emulsifications of toluene derivatives at 90‡C, the time that turbid water layers of surfactant solutions take to become clear is the same as that of the catalytic effect on oxidation of toluene derivatives. Thus, it can be inferred that surfactants can improve the oxidation yields of toluene derivatives by increasing the contact between two reacting phases.  相似文献   

15.
The effects of reaction conditions and concentration of the heteropoly acid H4PMo11VO40 supported onto ShAS-2 bead aluminosilicate on the conversion of toluene into benzoic acid in the partial oxidation of toluene by atmospheric oxygen were studied. The results demonstrated that the conversion of toluene was an extremal function of temperature, space velocity (v), and toluene concentration (C 0) in the initial air mixture. An increase in the heteropoly acid concentration from 2 to 30% increased the conversion of toluene into benzoic acid in the partial oxidation of toluene from 1.5 to 12.6% at optimum process parameters: T = 300°C, v = 2000 h-1, and C 0 = 13.72 g/m3.  相似文献   

16.
The kinetics of the reactions between sodium nitrite and phenol or m-, o-, or p-cresol in potassium hydrogen phthalate buffers of pH 2.5–5.7 were determined by integration of the monitored absorbance of the C-nitroso reaction products. At pH > 3, the dominant reaction was C-nitrosation through a mechanism that appears to consist of a diffusion-controlled attack on the nitrosatable substrate by NO+/NO2H2+ ions followed by a slow proton transfer step; the latter step is supported by the observation of basic catalysis by the buffer which does not form alternative nitrosating agents as nitrosyl compounds. The catalytic coefficients of both anionic forms of the buffer have been determined. The observed order of substrate reactivities (o-cresol ≈ m-cresol > phenol ≫ p-cresol) is explained by the hyperconjugative effect of the methyl group in o- and m-cresol, and by its blocking the para position in p-cresol. Analysis of a plot of ΔH# against ΔS# shows that the reaction with p-cresol differs from those with o- and m-cresol as regards the formation and decomposition of the transition state. The genotoxicity of nitrosatable phenols is compared with their reactivity with NO+/NO2H2+. © 1997 John Wiley & Sons, Inc.  相似文献   

17.
1-, 2-cis-, 2-trans-, and 3-trans-heptenes (C7)are isomerized either very slowly or not at all with IrX(CO)L2 at 80°C in toluene and under N2. However, under the conditions of hydrogenation fast isomerisation takes place. With IrCl(CO)L2 as catalyst the rate of isomerisation decreases the order: 1-C7 ∼ 2-cis-C7 > 3-trans-C7 > 2-trans-C7. This sequence is independent of the ligand L in lrCl(CO)L2, however, with a particular isomer the rate of isomerisation is a function of L in the order L = PPh3 > P(C6H11)3 > P(OPh)3.  相似文献   

18.
Wenbin Zeng  Dieter Hoppe 《Tetrahedron》2005,61(13):3281-3287
(E)/(Z)-Isomeric allylic carbamate esters were deprotonated by n-butyllithium/(−)-sparteine in toluene. Trapping experiments with chlorotrimethylsilane afforded the α-substitution products, with (R)-configuration, revealing that the pro-S proton is removed predominantly to form the corresponding (S)-lithium·(−)-sparteine derivatives; kS/kR>15:1 and >7:1, respectively. A slow (S)→(R)-epimerization occurs at −78 °C (T1/2>60 min). The allylic double bond is stable to (Z)-(E) isomerization under these conditions.  相似文献   

19.
A reference material for the biological monitoring of occupational exposure to toluene, benzene and phenol was prepared. O-cresol and hippuric acid (metabolites of toluene) are used for the biological monitoring of occupational exposure to toluene. Phenol, a metabolite of benzene, is used for the biological monitoring of exposure to benzene, but phenol can of course also be used as an indicator of exposure to phenol as well. The reference material (RM) used for the determination of these metabolites was prepared by freeze-drying pooled urine samples obtained from healthy persons occupationally exposed to toluene and those taking part in an inhalation experiment. Tests for homogeneity and stability were performed by determining urine concentrations of o-cresol, hippuric acid, creatinine and phenol. To investigate the stability of the RM, the urinary concentrations of o-cresol and phenol were monitored for eighteen months using GC and HPLC, while those of hippuric acid and creatinine were followed for five and six years, respectively, using HPLC. Analysis of variance showed that the concentrations did not change. The certified concentration values (and their uncertainties) of the substances in this reference material (phenol concentration c=6.46±0.58 mg l−1; o-cresol concentration c=1.17±0.15 mg l−1; hippuric acid concentration c=1328±30 mg l−1; creatinine concentration c=0.82±0.10 g l−1) were evaluated via the interactive statistical programme IPECA.  相似文献   

20.
Iodine is shown to be an efficient catalyst for a one-step, three-component aza-Friedel–Crafts reaction of activated arenes or heteroarenes with benzyl or tert-butyl carbamates in combination with a wide variety of aldehydes in toluene under ‘open-flask’ and mild conditions. In the presence of 5 mol % of iodine in toluene at room temperature, the reaction gives the corresponding N-CBz or N-Boc protected α-branched amines, selectively, in good to excellent yields.  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