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.
OH radicals, produced in aqueous acid media by the electrogenerated system Fe(II)/H2O2, promote the oxidation of polymethylated benzenes. Under our experimental conditions only one -CH3 group was oxidized to -CHO. A reduction of aldehyde to alcohol by.OOH radicals was shown to occur when H2O2 was produced in excess over Fe(II). In hydrochloric acid or in perchloric acid with Cl ions added, a remarkable increase of the yields was observed. The results are discussed on the basis of the kinetic parameters of the reactions involved in the process.  相似文献   

2.
Iron oxide and TiO2 were immobilized on modified polyvinyl fluoride films in a sequential process. Synergic effects of iron oxide and TiO2 on the polymer film were observed during the heterogeneous degradation of hydroquinone (HQ) in the presence of H2O2 at pH close to neutrality and under simulated solar irradiation. Within the degradation period, little iron leaching (<0.5 mg/L) was observed.The surface of commercial polyvinyl fluoride (PVF) film was modified by TiO2 under light inducing oxygen group (C–OH, CO, COOH) formation on the film surface. During this treatment, TiO2 nanoparticles simultaneously bind to the film, leading to PVFf–TiO2. The possible mechanistic pathway for the TiO2 deposition and the nature of the polymer–TiO2 interaction are discussed. Furthermore PVF and PVFf–TiO2 were immersed in an aqueous solution for the deposition of iron oxide layer by hydrolysis of FeCl3, leading to PVF–Fe oxide and to PVFf–TiO2–Fe oxide respectively.HQ degradation and mineralization mediated by PVFf–TiO2, PVF–Fe oxide and PVFf–TiO2–Fe oxide were investigated under different conditions. Remarkable synergistic effects were observed for PVFf–TiO2–Fe oxide possibly due to Fe(II) regeneration, accelerated by electron transfer from TiO2 to the iron oxide under light.  相似文献   

3.
The reaction pathways for NH2+O2Products are considered on the basis of experimental data on the ignition of an NH3/O2/Ar mixture in reflected shock waves (p=1–10 atm, T=900–2160 K) and on the NH3/O2/Ar flame structure (p=35 torr, T=1050–2600 K) using a multistage kinetic mechanism. The rate constants of the NH2+O2=HNO+OH reaction, obtained from a comparison of experimental and calculated data, are reported (k=3·1011 exp (–15,000/RT) cm3/(mole·s) at T1500–2160 L and k=3·109 cm3/(mole·s) at T900–1400 K).Novosibirsk. Translated from Fizika Goreniya i Vzryva, Vol. 30, No. 1, pp. 60–65, January–February, 1994.  相似文献   

4.
The effect of benzotriazole, BTA, on mass transfer in dissolution-corrosion of the copper rotating disk electrode in 0.02 M Fe(III)–0.5 M H2SO4 has been studied by means of atomic absorption spectrometry. The mass transfer coefficient, K, was determined from the slope of ln(C 0/C)Fe(III) vs. time plots. In the absence of BTA the corrosion process can be described by the correlation Sh = KR/D = 4.47Re 0.5. The difference in values between Sh and Sh Levich, and the change in slope in the Arrenhius plot points to mixed control for the cathodic process Fe3+ + 1e Fe 2+ and charge transfer control for the anodic process, Cu Cu2+ + 2e. The average activation energies were 7.7 kJ mol–1 and 19.5 kJ mol–1 at (500–1500) and (2000–3000) rpm, respectively. At low concentration of BTA the inhibiting action of BTA increases with concentration and with rotation speed. For [BTA] 5 × 10–3 M, the K value, 10–4 cm s–1, remains constant and is independent of rotation rate. The morphology of the copper rotating disk after corrosion in the absence and presence of BTA was examined using scanning electron microscopy (SEM).  相似文献   

5.
Two methods were used to remove Cr(VI) from industrial wastewater. Although both are based in the same general reaction: 3Fe(II)(aq) + Cr(VI)(aq) ; 3Fe(III)(aq) + Cr(III)(aq) the way in which the required amount of Fe(II) is added to the wastewater is different for each method. In the chemical method, Fe(II)(aq) is supplied by dissolving FeSO4 · 7(H2O)(s) into the wastewater, while in the electrochemical process Fe(II)(aq) ions are formed directly in solution by anodic dissolution of an steel electrode. After this reduction process, the resulting Cr(III)(aq) and Fe(III)(aq) ions are precipitated as insoluble hydroxide species, in both cases, changing the pH (i.e., adding Ca(OH)2(s)). Based on the chemical and thermodynamic characteristics of the systems Cr(VI)–Cr(III)–H2O–e and Fe(III)–Fe(II)–H2O–e both processes were optimized. However we show that the electrochemical option, apart from providing a better form of control, generates significantly less sludge as compared with the chemical process. Furthermore, it is also shown that sludge ageing promotes the formation of soluble polynuclear species of Cr(III). Therefore, it is recommended to separate the chromium and iron-bearing phases once they are formed. We propose the optimum hydraulic conditions for the continuous reduction of Cr(VI) present in the aqueous media treated in a plug-flow reactor.  相似文献   

6.
Extended studies on Zn-ZSM-5 catalyst for the production of liquid hydrocarbons in the direct partial oxidation (DPO) of CH4 with O2 are reported. Previously, it was reported that metal-containing ZSM-5 catalysts could produce C5+ hydrocarbons from pure CH4/O2 feeds without feed additives. Zn-ZSM-5 produced the highest C5+ yields of the catalysts tested. This work shows that the method of introducing Zn onto the catalyst, ion-exchange versus impregnation, does not significantly alter C5+ yields if low Zn content is maintained ( 0.4–0.5 wt%). Liquid hydrocarbon yields in this system doubled after 8 h on stream while overall C2+ yields increased by over 300%. Mechanistic implications of these findings are discussed. Finally, processing a natural gas feed over Zn-ZSM-5 gave higher C5+ yields over CH4 feed but these yields were not improved over previously published results using HZSM-5.  相似文献   

7.
Production of hydroxyl radical (OH) is of significant concern in engineered and natural environment. A simple in situ method was developed to measure OH formation in UV/H2O2, UV/Fe(III), and UV/NO3? systems using trapping of OH by benzoic acid (BA) and measuring fluorescence signals from hydroxylated products of BA. Method development included characterization of OH trapping mechanism and measurement of quantum yields (ΦOH) for OH. The distribution of OHBA isomers was in the order of o-OHBA > p-OHBA > m-OHBA, although it changed with the H2O2 concentration and light intensity. This supports that OH attacks dominantly on the benzene rings. The quantum yields for OH formation in the UV/H2O2 process were 1.02 and 0.59 at 254 and 313 nm, which were in good agreement with the literature values, confirming that the method is suitable for the measurement of OH production from UV/H2O2 processes. Using the continuous flow method developed, quantum yields for OH in UV/H2O2, UV/Fe(III), and UV/NO3? systems were measured varying the initial concentration of OH precursors. The ΦOH values increased with increasing concentrations of H2O2, Fe(III), and NO3? and approached constant values as the concentration increased. The ΦOH values were 0.009 for H2O2 at 365 nm, showing that OH production is not negligible at such high wavelength. The ΦOH values during the photolysis of Fe(OH)2+ (pH 3.0) and Fe(OH)2+ (pH 6.0) at 254 nm were 0.34 and 0.037, respectively. The ΦOH values for NO3? approached a constant value of 0.045 at 254 nm at the initial concentration of 10 mM.  相似文献   

8.
The addition of F to Ba-Ti mixed oxide catalysts significantly improves the catalytic performances for the oxidative coupling of methane (MOC), which can achieve high C2 yields at wide feed composition range and high GHSV. The effect is particularly marked for the BaF2– TiO2 catalysts containing more than 50 mol% BaF2. The C2 yield of 17% and the C2 selectivity of > 60% were achieved over these catalysts at 700 ° C. After being on stream for 31 h, the 50 mol% BaF2-TiO2 catalysts showed only a 1–1.5% decrease in the C2 yields. Results obtained by XRD show that various Ba-Ti oxyfluoride phases were formed due to the substitution of F to O2–.  相似文献   

9.
Solutions of the veterinary fluoroquinolone antibiotic enrofloxacin in 0.05 M Na2SO4 of pH 3.0 have been comparatively degraded by electrochemical advanced oxidation processes such as anodic oxidation with electrogenerated H2O2 (AO-H2O2), electro-Fenton (EF), photoelectro-Fenton (PEF) and solar photoelectro-Fenton (SPEF) at constant current density. The study has been performed using an undivided stirred tank reactor of 100 ml and a batch recirculation flow plant of 2.5 l with an undivided filter-press cell coupled to a solar photoreactor, both equipped with a Pt or boron-doped diamond (BDD) anode and a carbon-polytetrafluoroethylene gas diffusion cathode to generate H2O2 from O2 reduction. In EF, PEF and SPEF, hydroxyl radical (OH) is formed from Fenton's reaction between added catalytic Fe2+ and generated H2O2. Almost total decontamination of enrofloxacin solutions is achieved in the stirred tank reactor by SPEF with BDD. The use of the batch recirculation flow plant showed that this process is the most efficient and can be viable for industrial application, becoming more economic and yielding higher mineralization degree with raising antibiotic content. This is feasible because organics are quickly oxidized with OH formed from Fenton's reaction and at BDD from water oxidation, combined with the fast photolysis of complexes of Fe(III) with generated carboxylic acids under solar irradiation. The lower intensity of UVA irradiation used in PEF with BDD causes a slower degradation. EF with BDD is less efficient since OH cannot destroy the most persistent Fe(III)-oxalate and Fe(III)-oxamate complexes. AO-H2O2 with BDD yields the poorest mineralization because pollutants are only removed with OH generated at BDD. All procedures are less potent using Pt as anode due to the lower production of OH at its surface. Enrofloxacin decay always follows a pseudo first-order reaction. Its primary aromatic by-products and short intermediates including polyols, ketones, carboxylic acids and N-derivatives are detected by GC-MS and chromatographic techniques. The evolution of F, NO3 and NH4+ ions released to the medium during each process is also determined.  相似文献   

10.
The structural role of copper ions in melts (glasses) of the Na2O–SiO2–Cu2O–CuO system is analyzed in the framework of the acid–base concept with due regard for the geometric (the radius ratio for Cu2(1)+ and O2– ions) and energy (the mean enthalpies of the Cu2(1)+–O bonds) factors. It is demonstrated that copper ions in the structure fulfill the function of modifier cations. In these melts, the Cu1+–Cu2+ redox equilibrium can be described without regard for the formation of [Cu2(1)+O4/2]2(3)– ionic complexes (which could be incorporated into the structure of silicon–oxygen anions) and [Cu2+O b/k ]2 – b/k polyhedra providing the interaction between Cu2+ ions and anions. The influence of the formation of these polyhedra on the redox equilibrium is considered within the formalism of chemical thermodynamics. The composition dependence of the oxygen ion exponent pO is measured by an electromotive force (emf) technique. The ratio between the numbers of copper atoms with different valences is determined by chemical analysis. The experimental data obtained are in agreement with the theoretical inferences.  相似文献   

11.
The production of Co(III) acetate from Co(II) acetate using a bipolar trickle tower of graphite Raschig rings was investigated. Space time yields up to 18 kg m–3 h–1 were obtained, which showed no improvement over those achievable in a conventional plate and frame cell. A mathematical model of the system indicated that the electrode reactions occurred almost entirely at the opposing annular surfaces between consecutive layers of Raschig rings and that the unexpectedly low performance of the device was most probably due to the unfavourable mass transport conditions which existed in the intervening gaps.Nomenclature a annular cross sectional area of one Raschig ring (m2) - b C kinetic exponential constant for reduction of Co(III) (V–1) - b A kinetic exponential constant for oxidation of Co(II) (V–1) - b H kinetic exponential constant for hydrogen evolution (V–1) - b 0 kinetic exponential constant for oxygen evolution (V–1) - [Co(II)] concentration of Co(II) (mol m–3) - [Co(III)] concentration of Co(III) (mol m–3) - F Faraday constant (96 487 C mol–1) - f fraction of total flow by-passing the annular gap between adjacent Raschig rings in a vertical row - I current per vertical column of rings (A) - k C rate constant for reduction of Co(III) (A m mol–1) - k A rate constant for oxidation of Co(II) (A m mol–1) - k H rate constant for hydrogen evolution (A m–2) - k O rate constant for oxygen evolution (A m–2) - k L mass transfer coefficient (m s–1) - Q flow rate per vertical row of Raschig rings (m3s–1) - v volume of annular gap between adjacent Raschig rings in a vertical row (m3) - V superficial velocity of electrolyte (m s–1) - A anodic potential (V) - C cathodic potential (V)  相似文献   

12.
13.
The electrochemistry of molten LiOH–NaOH, LiOH–KOH, and NaOH–KOH was investigated using platinum, palladium, nickel, silver, aluminum and other electrodes. The fast kinetics of the Ag+/Ag electrode reaction suggests its use as a reference electrode in molten hydroxides. The key equilibrium reaction in each of these melts is 2 OH = H2O + O2– where H2O is the Lux-Flood acid (oxide ion acceptor) and O2– is the Lux–Flood base. This reaction dictates the minimum H2O content attainable in the melt. Extensive heating at 500 °C simply converts more of the alkali metal hydroxide into the corresponding oxide, that is, Li2O, Na2O or K2O. Thermodynamic calculations suggest that Li2O acts as a Lux–Flood acid in molten NaOH–KOH via the dissolution reaction Li2O(s) + 2 OH = 2 LiO + H2O whereas Na2O acts as a Lux–Flood base, Na2O(s) = 2 Na+ + O2–. The dominant limiting anodic reaction on platinum in all three melts is the oxidation of OH to yield oxygen, that is 2 OH 1/2 O2 + H2O + 2 e. The limiting cathodic reaction in these melts is the reduction of water in acidic melts ([H2O] [O2–]) and the reduction of Na+ or K+ in basic melts. The direct reduction of OH to hydrogen and O2– is thermodynamically impossible in molten hydroxides. The electrostability window for thermal battery applications in molten hydroxides at 250–300 °C is 1.5 V in acidic melts and 2.5 V in basic melts. The use of aluminum substrates could possibly extend this window to 3 V or higher. Preliminary tests of the Li–Fe (LAN) anode in molten LiOH–KOH and NaOH–KOH show that this anode is not stable in these melts at acidic conditions. The presence of superoxide ions in these acidic melts likely contributes to this instability of lithium anodes. Thermal battery development using molten hydroxides will likely require less active anode materials such as Li–Al alloys or the use of more basic melts. It is well established that sodium metal is both soluble and stable in basic NaOH–KOH melts and has been used as a reference electrode for this system.  相似文献   

14.
A numerical model has been developed to describe the behaviour of a batch reactor in which Fenton's reagent is used for hydroxylating aromatic hydrocarbons under conditions of electrochemical regeneration. The test reaction considered is the conversion of benzene into phenol. Comparison is made with previously published experimental results.Nomenclature A electrode area, m2 - a 1 parameter defined by Equation 21 - C i concentration of species, i, in the bulk solution, mol m–3 - c i local concentration of species, i, in the diffusion layer, mol m–3 - K i effective mass-transfer coefficient, m s–1 - k j rate constant of reaction j - R j rate of reaction j, mol m–3 s–1 - r i rate of change of concentration of species i due to chemical reaction, mol m–3 s–1 - t time, s - V reactor volume, m3 - x distance from the cathode surface, m - x * maximum thickness of the diffusion layer, m - period of diffusion layer renewal, s Subscrpts 1 oxygen - 2 Fe3+ - 3 hydrogen peroxide - 4 Fe2+ - 5 benzene - 6 phenol - 7 biphenyl This paper was presented at the meeting on Electroorganic Process Engineering held in Perpignan, France, 19–20 September 1985.  相似文献   

15.
This paper reports a comparative study on the anodic oxidation of 2.5 l of 50 mg l−1 TOC of formic, oxalic, acetic, pyruvic or maleic acid in 0.1 M Na2SO4 solutions of pH 3.0 with and without 1.0 mM Fe3+ as catalyst in the dark or under solar irradiation. Experiments have been performed with a batch recirculation flow plant containing a one-compartment filter-press electrolytic reactor equipped with a 20 cm2 boron-doped diamond (BDD) anode and a 20 cm2 stainless steel cathode, and coupled to a solar photoreactor. This system gradually accumulates H2O2 from dimerization of hydroxyl radical (OH) formed at the anode surface from water oxidation. Carboxylic acids in direct anodic oxidation are mainly oxidized by direct charge transfer and/or OH produced on BDD, while their Fe(III) complexes formed in presence of Fe3+ can also react with OH produced from Fenton reaction between regenerated Fe2+ with electrosynthesized H2O2 and/or photo-Fenton reaction. Fast photolysis of Fe(III)-oxalate and Fe(III)-pyruvate complexes under the action of sunlight also takes place. Chemical and photochemical trials of the same solutions have been made to better clarify the role of the different catalysts. Solar photoassisted anodic oxidation in presence of Fe3+ strongly accelerates the removal of all carboxylic acids in comparison with direct anodic oxidation, except for acetic acid that is removed at similar rate in both cases. This novel electrochemical advanced oxidation process allows more rapid mineralization of formic, oxalic and maleic acids, without any significant effect on the conversion of acetic acid into CO2. The synergistic action of Fe3+ and sunlight in anodic oxidation can then be useful for wastewater remediation when oxalic and formic acids are formed as ultimate carboxylic acids of organic pollutants, but its performance is expected to strongly decay in the case of generation of persistent acetic acid during the degradation process.  相似文献   

16.
Diffusion combustion of ethanol and hydrogen and homogeneous combustion of hydrogen–oxygen mixtures are studied by the laserinduced fluorescence technique in the linear regime and with signal saturation. Data on the flame temperature and OH concentration are obtained. The burning temperature of 3090 K for a stoichiometric O2–H2 mixture is in agreement with the known value. It is shown that the maximum concentrations of radicals in the hydrogen–air and stoichiometric hydrogen–oxygen flames are close to each other (4.4· 1016 cm-3).  相似文献   

17.
Partial oxidation of methane by oxygen to form formaldehyde, carbon oxides, and C2 products (ethane and ethene) has been studied over silica catalyst supports (fumed Cabosil and Grace 636 silica gel) in the 630–780 °C temperature range under ambient pressure. The silica catalysts exhibit high space time yields (at low conversions) for methane partial oxidation to formaldehyde, and the C2 hydrocarbons were found to be parallel products with formaldehyde. Short residence times enhanced both the C2 hydrocarbons and formaldehyde selectivities over the carbon oxides even within the differential reactor regime at 780 °C. This suggests that the formaldehyde did not originate from methyl radicals, but rather from methoxy complexes formed upon the direct chemisorption of methane at the silica surface at high temperature. Very high formaldehyde space time yields (e.g., 812 g/kg cat h at the gas hourly space velocity = 560 000 (NTP)/kg cat h) could be obtained over the silica gel catalyst at 780 °C with a methane/air mixture of 1.5/1. These yields greatly surpass those reported for silicas earlier, as well as those over many other catalysts. Low CO2 yields were observed under these reaction conditions, and the selectivities to formaldehyde and C2 hydrocarbons were 28.0 and 38.8%, respectively, at a methane conversion of 0.7%. A reaction mechanism was proposed for the methane activation over the silica surface based on the present studies, which can explain the product distribution patterns (specifically the parallel formation of formaldehyde and C2 hydrocarbons).  相似文献   

18.
An electrochemical ozone generation process was studied wherein glassy carbon anodes and air depolarized cathodes were used to produce ozone at concentrations much higher than those obtainable by conventional oxygen-fed corona discharge generators. A mathematical model of the build up of ozone concentration with time is presented and compared to experimental data. Products based on this technology show promise of decreased initial costs compared with corona discharge ozone generation; however, energy consumption per kg ozone is greater. Recent developments in the literature are reviewed.Nomenclature A electrode area (m2) - Ar * modified Archimedes number, d b 3 gG/2 (1 — G) - C O 3 (aq) concentration of dissolved ozone (mol m–3) - C O 3 i concentration at interface (mol m–3) - C O 3 1 concentration in bulk liquid (mol m–3) - D diffusion coefficient (m2 s–1) - E electrode potential against reference (V) - F charge of one mole of electrons (96 485 C mol–1) - g gravitational acceleration (9.806 65 m s–2) - i current density (A m–2) - i 1 limiting current density (A m–2) - I current (A) - j material flux per unit area (mol m–2 s–1) - k obs observed rate constant (mol–1 s–1) - k t thermal conductivity (J s–1 K–1) - L reactor/anode height (m) - N O 3 average rate of mass transfer (mol m–2 s–1) - Q heat flux (J s–1) - r i radius of anode interior (m) - r a radius of anode exterior (m) - r c radius of cathode (m) - R gas constant (8.314 J K–1 mol–1) - S c Schmidt number, v/D - Sh Sherwood number, k m d b/D = i L d b/zFD[O3] - t time (s) - T i temperature of inner surface (K) - T o temperature of outer surface (K) - U reactor terminal voltage (V) - electrolyte linear velocity (m s–1) - V volume (m3) - V O 3 volume of ozone evolved (10–6 m3 h–1) - z i number of Faradays per mole of reactant in the electrochemical reaction Greek symbols G gas phase fraction in the electrolyte - (mean) Nernst diffusion layer thickness (m) - fractional current efficiency - overpotential (V) - electrolyte kinematic viscosity (m2 s–1) - electrolyte resistivity (V A–1 m)  相似文献   

19.
Cyclic voltammetric and potentiodynamic studies were carried out on 300W carbon steel in Bayer plant solution, at 100 °C, with different alumina concentrations. Alumina behaves as an anodic inhibitor, shifting the critical passivation potentials positively and decreasing the critical passivation current with increasing concentration. Increase in alumina concentration promotes the formation of a uniform and less porous film. The pore resistance model describes the properties of the oxide films. Aluminium was found in all oxides formed, supporting the formation of a mixed oxide Fe3–x Al x O4. Thermodynamic calculation of some equilibrium potentials was carried out using the Fe(OH)3 ion rather than HFeO2 ion. Moreover, the Al(OH)4 ion was considered instead of AlO2 ion in the oxidation process.  相似文献   

20.
Synthesis gas formation by direct oxidation of methane over Rh monoliths   总被引:7,自引:0,他引:7  
The production of H2 and CO by catalytic partial oxidation of CH4 in air or O2 at atmospheric pressure has been examined over Rh-coated monoliths at residence times between 10–4 and 10–2 s and compared to previously reported results for Pt-coated monoliths. Using O2, selectivities for H2 ( ) as high as 90% and CO selectivities (S CO) of 96% can be obtained with Rh catalysts. With room temperature feeds using air, Rh catalysts give of about 70% compared to only about 40% for Pt catalysts. The optimal selectivities for either Pt or Rh can be improved by increasing the adiabatic reaction temperature by preheating the reactant gases or using O2 instead of air. The superiority of Rh over Pt for H2 generation can be explained by a methane pyrolysis surface reaction mechanism of oxidation at high temperatures on these noble metals. Because of the higher activation energy for OH formation on Rh (20 kcal/mol) than on Pt (2.5 kcal/mol), H adatoms are more likely to combine and desorb as H2 than on Pt, on which the O+ H OH reaction is much faster.This research was partially supported by DOE under Grant No. DE-FG02-88ER13878-AO2.  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