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 concerted effects of hydroxyl and methyl substituents in controlling the site of .OH radical attack on aromatics in aqueous solutions are explored using the cresols as typical examples. The distributions of dihydroxytoluenes produced in the radiolysis of aqueous solutions of the cresols containing ferricyanide as a radical oxidant were examined by capillary electrophoretic and liquid chromatographic methods. Because .OH is a strong electrophile, it adds preferentially at the electron-rich sites of an aromatic ring. As a result, the observed distributions of dihydroxytoluenes reflect the charge distributions in the cresols. It is shown that in the case of m-cresol the hydroxyl substituent has a dominant ortho-para directing effect similar to that observed for phenol. In o- and p-cresol, this effect is modified, indicating that the methyl substituent has a significant effect on the electronic structure of those cresols. Correlation of the charge distribution in the cresols indicated by the observed distribution of dihydroxytoluenes with the unpaired spin distribution in the corresponding methylphenoxyl radicals demonstrates that the electronic structures of o- and p-cresol and their corresponding phenoxyl radicals are similarly affected by hydroxyl and methyl substitution. Addition of .OH at the methyl-substituted positions of o- and p-cresol to produce o- and p-dienone is also reported. The observation of these dienones demonstrates that addition of .OH at the ipso positions of alkylated aromatics can be of considerable importance. Mass spectrometric studies show that these dienones have relatively higher proton affinities than their isomeric analogues.  相似文献   

2.
Thermal reactivities of lignin pyrolysis intermediates, catechols/pyrogallols (O-CH3 homolysis products) and cresols/xylenols (OCH3 rearrangement products), were studied in a closed ampoule reactor (N2/600 °C/40-600 s) to understand their roles in the secondary reactions step. Reactivity tends to be enhanced by increasing the number of substituent groups on phenol and this effect was greater for -OH than for -CH3. Thus, catechols/pyrogallols were more reactive than cresols/xylenols and syringol-derived products were more reactive than corresponding guaiacol-derived products. Catechols/pyrogallols were effectively converted into CO (additionally CO2 in the case of pyrogallols) in the early stage of pyrolysis. In contrast, cresols/xylenols were comparatively stable and produced H2, CH4 and demethylation products (cresols and phenol) after prolonged heating. All intermediates except phenol and 2-ethylphenol formed coke during a long heating time of 600 s (second stage coking). Based on the present results, the roles of intermediates in tar, coke and gas formation from guaiacol and syringol are discussed at the molecular level, focusing on their differences. Molecular mechanisms of gas formation from pyrogallols and demethylation of cresols/xylenols are also discussed.  相似文献   

3.
The degradation of organic compounds in an aqueous solution induced by a gaseous plasma, which was sustained locally on the surface of solution by means of contact glow discharge electrolysis, was mechanistically studied. The oxidation of phenol was extensively examined and it was revealed that the rate of disappearance followed the first-order rate law when the initial concentration of phenol was lower than 100 mM. As the initial concentration increased, the rate equation gradually deviated from the first-order and eventually shifted to zero-order above 250 mM. Such a kinetical behavior could be rationalized by assuming the reaction scheme where hydroxyl radical would act as the most responsible key-species for the organic degradation in competition with the coupling to hydrogen peroxide. In addition, the catalytic effects of Fe ions in the solution on the rate of organic degradation were closely investigated and could be consistently explained on the above concept.  相似文献   

4.
The photocatalytic degradation of the antibiotic sulfamethazine under excitation at 365 nm of Pd‐doped ceria‐ZnO nanocomposite, titanium dioxide and iron(III) aqua complex was deeply studied from the analytical point of view. It reveals the formation of nine degradation products that were detected in their protonated forms using LC/electrospray ionization quadrupole time‐of‐flight MS in the positive mode. Their formation involves the hydroxyl radical, and their concentrations increased with irradiation time. Collision‐induced dissociation tandem mass spectrometry associated with the accurate mass measurements was efficiently used for the elucidation of their chemical structures. None of these identified degradation products has been already reported in the literature. Three by‐products result from the hydroxylation at the pyrimidine moiety as well as at the aromatic part, two of them arise from the scission of the pyrimidine group, and finally, three of them come from the scission of the sulfamide bridge. This points the evidence of studying the fate of these degradation products if their toxicity is demonstrated because they are clearly the result of the reaction of hydroxyl radical with the antibiotic sulfamethazine. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

5.
The reactions of three aromatic hydrocarbons, benzene, toluene, and styrene, with oxygen in a radiofrequency (r) plasma were investigated. Benzene was oxidized to yield phenol as a single volatile organic product. Similarly, toluene gave the ring oxidation products, cresols, as well as considerable amounts of methyl oxidation products, consisting mainly of benzaldehyde and benzyl alcohol. In contrast, the oxidation of styrene took place predominantly on the olefinic double bond to produce styrene oxide On basis of the products and effects of reaction variables, r power and flow rates of hydrocarbons and oxygen, on the reaction rate, the oxidation mechanism was discussed, particularly focusing on the intermediate species responsible for the formation of major products.  相似文献   

6.
The electrical decomposition of 4-chlorophenol in water was examined with iridium dioxide doped on atitanium electrode. A number of electrical degradation products of 4-chlorophenol, such as hydroquinone and chlorohydroquinone via the addition of hydroxyl radicals, and dichlorophenol through addition of chlorine radical, were observed as major products. Moreover, hydroxylated chlorobiphenylethers, hydroxylated dibenzo-p-dioxin/furans and hydroxylated chlorobiphenyls formed by a dimerization process during the electrolysis process of 4-chlorophenol were also observed. On the other hand, benzoquinone, muconic acid and aldehyde derivatives that were further oxidative products of hydroquinone formed by photocatalysis process, were not observed. The electrical decomposition products of 4-chlorophenol were trimethylsilylated and then identified by gas chromatography-mass spectrometry. The degradation rate of 4-chlorophenol in water by iridium oxide electrode was measured against the electrical process duration. After iridium electrical process for 120 min, about 50% of 4-chlorophenol was converted into a number of products through oxidation processes. On the basis of the identified products, the degradation pathways of 4-chlorophenol under electrolysis process were proposed.  相似文献   

7.
以Na2SO4为支持电解质, 使用Ti/PbO2电极, 研究了带有推电子基(—CH3)和吸电子基(—NO2, —Cl)的邻或对位取代基苯胺类化合物的电催化氧化降解过程. 研究结果表明, 带有取代基苯胺类化合物的氧化降解是在羟基自由基进攻下生成氨基酚类化合物, 然后在电极表面失去电子生成苯醌继续氧化的过程. 带有推电子基团苯胺的电催化降解速度比带有吸电子基团的苯胺降解速度快, 这是因为推电子基团使苯环电子云密度提高, 有利于羟基自由基的进攻; 吸电子基团使苯环电子云密度降低, 不利于羟基自由基的进攻. 由于阴极还原反应的作用, 化学反应活性和电化学反应活性并不完全一致. 氯代苯胺在羟基自由基进攻下—Cl离去, 以Cl-离子形式进入溶液中, 被氧化生成有效氯, 加快降解反应速度. 硝基虽然是强吸电子基, 但是可以转化为对苯二胺, 进一步活化苯环, 其降解速度较快.  相似文献   

8.
辉光放电等离子体处理阳离子染料结晶紫废水   总被引:2,自引:0,他引:2  
高锦章  马东平  郭晓  李岩  杨武 《应用化学》2007,24(5):534-539
用辉光放电等离子体技术对结晶紫进行了降解脱色处理,考察了多种因素对结晶紫降解效果的影响。实验发现,提高电解质浓度和增加电压均可提高结晶紫的脱色效果,考虑到电极损耗,辉光放电最佳条件为:电解质浓度为2 g/L Na2SO4,电压为600 V。当改变溶液的初始pH值时,结晶紫的脱色率随溶液的初始pH值升高而增加,加入一定量H2O2能明显地提高结晶紫的脱色效率;若加入0.4 mmol/L Fe2 ,5 min时结晶紫的脱色率由原来的13.64%增加到91.36%。结果表明,辉光放电产生的.OH对结晶紫的降解起重要作用。最佳条件下,40 min内的脱色率达到93%,降解率为74%。  相似文献   

9.
Pulsed Corona Discharge-Induced Reactions of Acetophenone in Water   总被引:2,自引:0,他引:2  
The reactions of acetophenone in water by pulsed corona discharges have been investigated to provide fundamental information concerning the reactions of acetophenone in water. Experimental results indicated that photolysis of acetophenone did not involve a hydroxyl radical mechanism and the majority flux of hydroxyl radicals originated from the dissociation of gas-phase oxygen in the plasma channels. The rate constants for photolysis and pyrolysis were determined to be 1.5×10–7 M-s–1, 2.2×10–4 s–1, respectively. The rate constant for the oxidative reactions was measured as 1.2×10–7 M-s–1. Results from this study support the proposal that acetophenone degradation reaction proceed through the oxidative reaction pathway, where molecular oxygen accelerates acetophenone degradation, photolysis, and pyrolysis pathways.  相似文献   

10.
A novel degradation study of prednisone (PRE) in aqueous systems (10 mg L?1) was performed under electron beam irradiation (EBI) in various conditions. The data demonstrate that the highest degradation caused by EBI was obtained with the oxidation of hydroxyl radicals. The proper amount of hydrogen peroxide could increase the PRE total carbon content removal rate. Nine degradation products of PRE in aqueous solution were preliminarily analyzed using ultra-high performance liquid chromatography in combination with mass spectrometry. Furthermore, the chemical analysis and possible degradation pathway were proposed with the assistance of theoretical calculation results by density functional theory.  相似文献   

11.
The mechanism of hydroxyl radical initiated degradation of a typical oil sands process water (OSPW) alicyclic carboxylic acid was studied using cyclohexanoic acid (CHA) as a model compound. By use of vacuum ultraviolet irradiation (VUV, 172 nm) and ultraviolet irradiation in the presence of hydrogen peroxide UV(254 nm)/H(2)O(2), it was established that CHA undergoes degradation through a peroxyl radical. In both processes the decay of the peroxyl radical leads predominantly to the formation of 4-oxo-CHA, and minor amounts of hydroxy-CHA (detected only in UV/H(2)O(2)). In UV/H(2)O(2), additional 4-oxo-CHA may also have been formed by direct reaction of the oxyl radical with H(2)O(2). The oxyl radical can be formed during decay of the peroxyl-CHA radical or reaction of hydroxy-CHA with hydroxyl radical. Oxo- and hydroxy-CHA further degraded to various dihydroxy-CHAs. Scission of the cyclohexane ring was also observed, on the basis of the observation of acyclic byproducts including heptadioic acid and various short-chain carboxylic acids. Overall, the hydroxyl radical induced degradation of CHA proceeded through several steps, involving more than one hydroxyl radical reaction, thus efficiency of the UV/H(2)O(2) reaction will depend on the rate of generation of hydroxyl radical throughout the process. In real applications to OSPW, concentrations of H(2)O(2) will need to be carefully optimized and the environmental fate and effects of the various degradation products of naphthenic acids considered.  相似文献   

12.
13.
The photolytic and photocatalytic degradation of the copolymers poly(methyl methacrylate-co-butyl methacrylate) (MMA-BMA), poly(methyl methacrylate-co-ethyl acrylate) (MMA-EA) and poly(methyl methacrylate-co-methacrylic acid) (MMA-MAA) have been carried out in solution in the presence of solution combustion synthesized TiO2 (CS TiO2) and commercial Degussa P-25 TiO2 (DP 25). The degradation rates of the copolymers were compared with the respective homopolymers. The copolymers and the homopolymers degraded randomly along the chain. The degradation rate was determined using continuous distribution kinetics. For all the polymers, CS TiO2 exhibited superior photo-activity compared to the uncatalysed and DP 25 systems, owing to its high surface hydroxyl content and high specific surface area. The time evolution of the hydroxyl and hydroperoxide stretching vibration in the Fourier transform-infrared (FT-IR) spectra of the copolymers indicated that the degradation rate follows the order MMA-MAA > MMA-EA > MMA-BMA. The same order is observed for the rate coefficients of photocatalytic degradation. The photodegradation rate coefficients were compared with the activation energy of pyrolytic degradation. In degradation by pyrolysis, it was observed that MMA-BMA was the least stable followed by MMA-EA and MMA-MAA. The observed contrast in the order of thermal stability compared to the photo-stability of these copolymers was attributed to the two different mechanisms governing the scission of the polymer and the evolution of the products.  相似文献   

14.
以商用TiO2P25为催化剂,分别在TiO2/UV/O2和TiO2/UV/N2两种体系下进行降解对氯硝基苯(pCNB)试验.采用ESR对两种体系下光催化反应形成的·OH进行测定,利用LC-MS对两种体系下反应形成的中间产物进行了定性和定量分析,最后对pCNB降解过程中氯和硝基的存在形式进行了研究.结果表明:TiO2/UV/O2体系的催化降解效果要明显优于TiO2/UV/N2体系;两种反应体系都有·OH产生,并且TiO2/UV/O2体系产生的·OH的量多于TiO2/UV/N2体系产生的·OH的量;TiO2/UV/O2体系形成的中间产物的种类要多于TiO2/UV/N2体系形成的,苯环上的氢、氯、硝基均可被·OH取代形成对硝基酚(pNP)、5-氯-2-硝基酚(5-C-2-PN)等酚类物质;两种体系下均有Cl-和NO2-存在,其中Cl-生成势与pCNB的去除势一致,只有TiO2/UV/O2体系中存在NO3-.  相似文献   

15.
The degradation of aqueous phase diphenyl induced by plasma that was locally generated between an electrolytic solution and an anode tip is investigated. Results indicated that the degradation rate can be greatly increased by increasing the applied voltage. Faster removal rate can be achieved for a relatively higher alkalinity or acidity. Fe2+ has an evident catalytic effect on the diphenyl elimination, while the presence of n‐butanol inhibited the degradation. The major intermediate products of the degradation process were identified by HPLC analysis and a degradation pathway was proposed.  相似文献   

16.
Liquid chromatography coupled with time-of-flight mass spectrometry (LC/TOFMS) was applied for the identification of four new photodegradation products of triclosan, a major antimicrobial agent used in personal care products. Wastewater samples, spiked at 7 microg/mL with triclosan, were irradiated with natural sunlight in order to generate the photodegradation products. Aliquots of the spiked water samples were taken at different times of irradiation and compounds were isolated from the water samples by solid-phase extraction. Separation and detection of the compounds and degradation products were accomplished by LC/TOFMS, which provided highly selective information about elemental compositions. Accurate mass measurements for the four degradation products permitted postulation of proposed empirical formulae in this study. Replacement of chlorine atoms by hydroxyl groups and chlorine losses are the major degradation pathways proposed. The degradation products were formed also under environmental conditions in wastewater matrices, thus suggesting their presence in real wastewater treatment processes.  相似文献   

17.
This study uses density functional theory (DFT) simulations to predict the main pathways by which hydroxyl (OH) radicals oxidize phenol into monohydroxylated products during an electrical discharge directly in or contacting water. The calculated activation energies and reaction rate constants indicate that phenol ring H abstraction is less likely to occur than OH addition, which will be the fastest in the ortho and para positions. The chain propagation with molecular oxygen of such formed ortho and para radicals will result in the production of hydroquinone and catechol, which are, concurrently, the most likely products of phenol degradation by OH radicals. Electron transfer reactions between dihydroxycyclohexadienyl radicals and plasma oxidative species are another important reaction mechanism which may be contributing significantly to the formation of products. Good agreement between computed kinetic and experimental data demonstrates the feasibility of applying DFT to investigate chemical reaction mechanisms.  相似文献   

18.
Ortiz CS  de Bertorello MM 《Talanta》1998,46(6):1537-1545
The goal of this study was to determine the kinetic parameters involved in the decomposition of 2-(5-methyl-4-isoxazolylamino)-N-(5-methyl-4-isoxazolyl)-1,4-naphthoquinone-4-imine (1) in aqueous solution and to identify the main degradation products. An isocratic HPLC assay was used to study the degradation rate of 1. The products of hydrolysis were identified by comparison of their retention times with those of authentic samples. The amount of 1 and the two degradation products resulting from storage of 1 in various buffer solutions was followed in function of time by a reversed-phase HPLC stability-indicating method. The observed degradation rates followed pseudo-first-order kinetics at constant pH, temperature and ionic strength. The log k–pH-profile was constructed at 35°C from the first-order rate constants obtained from studies at pH values ranging from 0.88 to 10.80 (μ=0.5 M). Hydrolysis in the acidic and alkaline media resulted in the formation of two degradation products in each case. The pH-rate profile of 1 in buffer solution was adequately described using a four-term rate equation. The obtained pH-rate profile indicated specific acid–base catalysis with a region of maximum stability between pH 6.40 and 7.40 which can be adequate for formulations of 1.  相似文献   

19.
Microwave-assisted photocatalytic (MPC) degradation of malachite green (MG) in aqueous TiO2 suspensions was investigated. A 20 mg/L sample of MG was rapidly and completely decomposed in 3 min with the corresponding TOC removal efficiency of about 85%. To gain insight into the degradation mechanism, both GC-MS and LC-ESI-MS/MS techniques were employed to identify the major intermediates of MG degradation, including N-demethylation intermediates [(p-dimethylaminophenyl)(p-methylaminophenyl)phenylmethylium (DM-PM), (p-methylaminophenyl)(p-methylaminophenyl)phenylmethylium (MM-PM), (p-methylaminophenyl)(p-aminophenyl)phenylmethylium (M-PM)]; a decomposition compound of the conjugated structure (4-dimethylaminobenzophenone (DLBP)); products resulting from the adduct reaction of hydroxyl radical; products of benzene removal; and other open-ring intermediates such as phenol, terephthalic acid, adipic acid, benzoic acid, etc. The possible degradation mechanism of MG included five processes: the N-demethylation process, adduct products of the hydroxyl radical, the breakdown of chromophores such as destruction of the conjugated structure intermediate, removal of benzene, and an open-ring reaction. To the best of our knowledge, it is the first time the whole MG photodegradation processes have been reported.  相似文献   

20.
Hydrolysis and methanolysis of Avilamycin C yielded a series of mono- and oligosaccharide-like products, which were, with one exception, identical with degradation products of flambamycin. Instead of evalose (6-deoxy-3-methyl-D -mannose), a building stone of Flambamycin and Everninomicin B, evermicose (2, 6-dideoxy-3-methyl-D -mannose) was identified as a constituent of the Avilamycins. This sugar was known as a degradation product of the Everninomicins C and D. From the degradation results a structure of Avilamycin A was postulated which differs from that of Flambamycin only by the lack of a hydroxyl group in position 2 of the evalose residue. This hypothesis was confirmed by a careful 1H- and 13C-NMR. study.  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