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 15 毫秒
1.
Photoformation rates and scavenging rate constants of hydroxyl radicals (OH) in natural water samples were determined by an automatic determination system. After addition of benzene as a chemical probe to a water sample in a reaction cell, light irradiation and injection of irradiated water samples into an HPLC as a function of time were performed automatically. Phenol produced by the reaction between OH and the benzene added to the water sample was determined to quantify the OH formation rate. The rate constants of OH formation from the photolysis of nitrate ions, nitrite ions and hydrogen peroxide were comparable with those obtained in previous studies. The percent of expected OH photoformation rate from added nitrate ion were high in drinking water (97.4%) and river water (99.3%). On the other hand, the low percent (65.0%) was observed in seawater due to the reaction of OH with the high concentrations of chloride and bromide ions. For the automatic system, the coefficient of variance for the determination of the OH formation rate was less than 5.0%, which is smaller than that in the previous report. When the complete time sequence of analytical cycle was 40 min for one sample, the detection limit of the photoformation rate and the sample throughput were 8 × 10−13 M s−1 and 20 samples per day, respectively. The automatic system successfully determined the photoformation rates and scavenging rate constants of OH in commercial drinking water and the major source and sink of OH were identified as nitrate and bicarbonate ions, respectively.  相似文献   

2.
The overall rate coefficient at standard temperature and pressure for the hydrogen abstraction reaction by the hydroxyl radical (HO˙) from common saturated volatile organic compounds (VOCs) is derived theoretically using electronic structure calculations and transition state theory (TST). The computational approach used is based on relatively efficient methods, and hence is applicable to a large number of compounds with only a modest use of computer resources. The key methods used are density functional theory (for the calculation of barrier heights) and simple transition state theory (TST), including a simple correction for tunnelling. All thermally relevant conformers of the reactant and the abstraction TS are included in the study. For all compounds in a test set of thirty-four, the calculated rate coefficient agrees with the experimental value to within better than an order of magnitude, and to within better than a factor of three for all but six cases, so that the accuracy is of predictive utility.  相似文献   

3.
4.
A hydrogen bonded complex between a hydroxyl radical and ozone has been found in argon matrices at 9 K. The shift of the OH stretch (-12.6 cm-1) indicates that this complex is somewhat weaker than the OH-CO complex (-21.8 cm-1, D0相似文献   

5.
The OH reaction with isoprene is studied. Methyl nitrite photolysis experiments were carried out in an outdoor smog chamber in an attempt to identify as completely as possible OH-isoprene product distribution. Emphasis was placed on identification and quantification of oxygenated products. A Tenax-based cryo-trap thermal desorber used to trap, concentrate, and dry chamber samples for identification on a GC/MS is described. Analysis of the products revealed that O(3P) can form in reaction systems designed to study OH reactions that include high concentrations of NO, and consequently NO2, hence, this reaction is also examined. The yields of methacrolein and methyl vinyl ketone are determined as 25 ± 3 and 35.5 ± 4%, respectively, with an additional 4 ± 2% as 3-methyl furan, totaling 65 ± 4%. These results, combined with those of previous studies allow 80% of isoprene's products to be explicitly identified, and the general structure of the remaining products to be ascertained. The O(3P) reaction produces 84 ± 8% epoxides, and 8 ± 3% species which result in production of HO2, and subsequently, OH. A heretofore unidentified product of the O(3P) reaction, 2-methyl 2-butenal, is identified. The rate constant of the NO2-isoprene reaction is measured.  相似文献   

6.
The reaction mechanism of carbonyl oxide with hydroxyl radical was investigated by using CASSCF, B3LYP, QCISD, CASPT2, and CCSD(T) theoretical approaches with the 6-311+G(d,p), 6-311+G(2df, 2p), and aug-cc-pVTZ basis sets. This reaction involves the formation of H2CO + HO2 radical in a process that is computed to be exothermic by 57 kcal/mol. However, the reaction mechanism is very complex and begins with the formation of a pre-reactive hydrogen-bonded complex and follows by the addition of HO radical to the carbon atom of H2COO, forming the intermediate peroxy-radical H2C(OO)OH before producing formaldehyde and hydroperoxy radical. Our calculations predict that both the pre-reactive hydrogen-bonded complex and the transition state of the addition process lie energetically below the enthalpy of the separate reactants (DeltaH(298K) = -6.1 and -2.5 kcal/mol, respectively) and the formation of the H2C(OO)OH adduct is exothermic by about 74 kcal/mol. Beyond this addition process, further reaction mechanisms have also been investigated, which involve the abstraction of a hydrogen of carbonyl oxide by HO radical, but the computed activation barriers suggest that they will not contribute to the gas-phase reaction of H2COO + HO.  相似文献   

7.
The rate constants and activation energies for the reactions of some thiophenes with the NO3 radical were measured using the absolute fast‐flow discharge technique at 263–335 K and low pressure. The proposed Arrhenius expressions for 2‐ethylthiophene, 2‐propylthiophene, 2,5‐dimethylthiophene, and 2‐chlorothiophene are k = (4.2 ± 0.28) ×10?16 exp[(2280 ± 70)]/T, k = (7.0 ± 2) × 10?18 exp[(3530 ± 70)]/T, k = (1 ± 1) × 10?14 exp[(1648 ± 240)]/T, and k = (8 ± 2) × 10?17 exp[(2000 ± 200)]/T (k = cm3 molecule?1 s?1), respectively. The reactions of this radical with 2‐chlorothiophene and 3‐chlorothiophene were also studied by a relative method in a Teflon static reactor at room temperature and atmospheric pressure. The effect of substitution on thiophene reactivity is discussed, and a relationship between the rate constants and the ionization potential (IP = ?EHOMO) has been proposed. © 2006 Wiley Periodicals, Inc. Int J Chem Kinet 38: 570–576, 2006  相似文献   

8.
Aqueous solutions containing the minichromosomal form of the virus SV40 and the radical scavenger DMSO were subjected to γ-irradiation, and the resulting formation of single-strand breaks (SSB) was quantified. Under the irradiation conditions, most SSBs were produced as a consequence of hydroxyl radical (OH) reactions. By controlling the competition between DMSO and the viral DNA substrate for OH, we are able to estimate the rate coefficient for the reaction of OH with the SV40 minichromosome. The results cannot be described adequately by homogeneous competition kinetics, but it is possible to describe the rate coefficient for the reaction as a function of the scavenging capacity of the solution. The experimentally determined rate coefficient lies in the range 1×109–2×109 L mol−1 s−1 at 107 s−1, and increases with increasing scavenging capacity.  相似文献   

9.
10.
Relative rate experiments were used to measure ratios of chemical kinetics rate constants as a function of temperature for the reactions of OH with isobutane, isopentane, 2-methylpentane, 3-methylpentane, 2,3-dimethylbutane, 2,3-dimethylpentane, 2,4-dimethylpentane, 2,3,4-trimethylpentane, n-heptane, n-octane, cyclopentane, cyclohexane, and cycloheptane. The results have been used to calibrate a structure-reactivity rate constant estimation method for k(298 K) which, when combined with previously determined relationships between k(298 K) and the Arrhenius parameters, is capable of determining the temperature dependence accurately. The estimation method reproduces most of the observed rate data within experimental accuracy but appears to fail for 2,3-dimethylbutane, which has an anomalously high rate constant. Curvature in the Arrhenius plots at low temperatures is not present for compounds with a single type of C-H bond and, for compounds with different C-H bonds, is shown to be consistent with effects due to different group sites on the molecule.  相似文献   

11.
In a study using UV photoelectron spectroscopy (PES) of the atmospherically relevant reaction
CH3SCH3 + Cl2 → CH3SCH2Cl + HCl
bands associated with a reaction intermediate have been observed. These have been assigned to ionization of the covalently bound molecule (CH3)2SCl2 on the basis of the intensity of the observed bands as a function of reaction time, molecular orbital calculations of vertical ionization energies and evidence from infrared spectroscopy.  相似文献   

12.
Various studies have been drawn toward the beneficial properties of fruit juices because they have several components, such as phenols, vitamins, and flavonoids, with antioxidant effects. However, fruit juices can also contain residues of pesticides used as standard pest control methods in crops. Many of these pesticides are degraded through oxidative mechanisms, and their persistence in juices can be enhanced by antioxidants. This study covers the degradation of four pesticides, aldicarb, demeton-S-methyl, fenamiphos, and methiocarb, to their respective sulfoxide and sulfone in grape juices, water (pH 3.5) and water (pH 3.5) with quercetin (one of the most important flavonoids of grape) added in an attempt to establish whether the presence of antioxidants can affect the degradation rate of pesticides. For this purpose, a multiresidue method based on solid-phase extraction (SPE) was developed for the simultaneous determination of these pesticides and their metabolites in commercial juices. The extraction procedure was carried out in C18 columns. The subsequent elution of pesticides was performed with dichloromethane prior to the determination by liquid chromatography-tandem mass spectrometry (LC-MS/MS), using two precursor-product ion transitions. Average recoveries for all the pesticides studied were higher than 80%, with relative standard deviations lower than 15% in the concentration range 0.005–0.05 μg/mL, and the quantification limits achieved ranged from 0.1 to 4.6 μg/L. The results demonstrated that degradation was slower in fruit juices and aqueous solutions with quercetin than in water. Several commercial grape juices were also analyzed to establish the levels of these pesticides. Methiocarb, fenamiphos, and demeton-S-methyl were found at low levels in some samples.  相似文献   

13.
The gas-phase reaction of n-butyl acetate with hydroxyl radicals has been studied in an environmental smog chamber at 298 K atmospheric pressure, and simulated tropospheric concentrations. The rate constant for this reaction has been determined by a relative method and the experimental result, relative to n-octane used as reference compound, is This value appears to be about 25% higher than absolute rate constants found in the literature, but agrees very well with the other relative determination. Two reaction products have been identified and their production yield has been estimated, each accounting for about (15 ± 5)% of the overall OH reaction processes. The two observed products are \centerline{ 2--oxobutil acetate ($\rm CH_3$--C0--0--$\rm CH_2$--CO--$\rm CH_2$--$\rm CH_3$)} and \centerline{ 2--oxobutil acetate ($\rm CH_3$--C0--0--$\rm CH_2$--$\rm CH_2$--CO--$\rm CH_3$)} The accuracy of the relative rate constant obtained is examined considering the evolution of the reactivity of the alkoxy end of the esters. Formation mechanisms for the two observed products are proposed and the likely other degradation channels are discussed. © 1996 John Wiley & Sons, Inc.  相似文献   

14.
Absolute rate constants and degradation efficiencies for hydroxyl radical and hydrated electron reactions with four different sulfa drugs in water have been evaluated using a combination of electron pulse radiolysis/absorption spectroscopy and steady-state radiolysis/high-performance liquid chromatography measurements. For sulfamethazine, sulfamethizole, sulfamethoxazole, and sulfamerazine, absolute rate constants for hydroxyl radical oxidation were determined as (8.3 +/- 0.8) x 10(9), (7.9 +/- 0.4) x 10(9), (8.5 +/- 0.3) x 10(9), and (7.8 +/- 0.3) x 10(9) M(-1) s(-1), respectively, with corresponding degradation efficiencies of 36% +/- 6%, 46% +/- 8%, 53% +/- 8%, and 35% +/- 5%. The reduction of these four compounds by their reaction with the hydrated electron occurred with rate constants of (2.4 +/- 0.1) x 10(10), (2.0 +/- 0.1) x 10(10), (1.0 +/- 0.03) x 10(10), and (2.0 +/- 0.1) x 10(10) M(-1) s(-1), respectively, with efficiencies of 0.5% +/- 4%, 61% +/- 9%, 71% +/- 10%, and 19% +/- 5%. We propose that hydroxyl radical adds predominantly to the sulfanilic acid ring of the different sulfa drugs based on similar hydroxyl radical rate constants and transient absorption spectra. In contrast, the variation in the rate constants for hydrated electrons with the sulfa drugs suggests the reaction occurs at different reaction sites, likely the different heterocyclic rings. The results of this study provide fundamental mechanistic parameters, hydroxyl radical and hydrated electron rate constants, and degradation efficiencies that are critical for the evaluation and implementation of advanced oxidation processes (AOPs).  相似文献   

15.
The kinetics of the reaction of OH radicals with methyl, n-propyl, and n-butyl nitrite have been studied in a discharge flow system under pseudo first-order conditions. The OH radicals were generated by the reaction of H atoms with NO2 and the concentration of OH; monitored by resonance fluorescence, was followed as a function of time in an excess of each nitrite. Values of k(CH3ONO) = (0.6 ± 0.09) × 109 dm3 mol?1 s?1 k(n – C3H7ONO) = (1.39 ± 0.20) × 109 dm3 mol?1 s?1, and k(n – C4H9ONO) = (2.89 ± 0.43) × 109 dm3 mol?1 s?1 at 295 K were obtained. These results agree with previous relative rate measurements from this laboratory but the value for k (CH3ONO) is a factor of 7 greater than the value obtained by relative rate measurements elsewhere using a different OH source.  相似文献   

16.
By using a simple treatment for the kinetics of radical polymerization with primary radical termination, the ratio kty/ktx of chain termination rate constant kty at conversion y to that ktx at conversion x and the ratio ktiy/ktix of the primary radical termination rate constant ktiy at conversion y to ktix at conversion x were calculated for the polymerizations of methyl methacrylate and ethyl acrylate in the conversion range 0 to 0.4. kty/ktx and ktiy/ktix were treated by using the following equations based on the variation of conversion: where g(T,y) is the average fractional free volume of radical chain end at conversion y and absolute temperature and β(T) is a function depending on T, and where gi(T,y) is the average fractional free volume of primary radical at conversion y and T and βi(T) is a function depending on T. The autoacceleration for the above monomers was successfully interpreted by the above treatment.  相似文献   

17.
18.
At low and high conversions, the chain termination rate constant for bimolecular termination between polymeric radicals given by kt = AtDs, where At is a constant and Ds is the diffusion constant of radical chain end, is completely correct. This termination rate constant does not depend on solution viscosity, but conversion.  相似文献   

19.
Methyl methacrylate and styrene were polymerized by using 2,2′-azobis(2,4-dimethyl valeronitrile) as initiator in various solvents. When a poor solvent is used, the dependence of polymerization rate Rp on initiator concentration [C] is small and can be treated by equations for the analysis of the polymerization with primary radical termination. With a good solvent, the dependence of Rp on [C] is so large that such equations are not applicable. Thus, the [C] dependence in a good solvent is explained qualitatively through the molecular weight dependence of rate for termination between polymer radicals, based on the excluded volume effect.  相似文献   

20.
Reactions of l-arginine (Arg) with hydroxyl radical (OH) and sulfate radical anion (SO4) were kinetically investigated by the pulse radiolysis technique. Hydrogen abstraction from Arg by OH afforded redox chemically oxidizing, neutral, and reducing carbon-centered Arg radicals. Kinetic properties of the radicals indicated that the reducing species might include the δ-C-centered Arg radical and CO2 radical anion. Similar transient spectra were observed in the SO4 reaction with Arg, suggesting direct oxidation at the guanidino group is less likely.  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