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 69 毫秒
1.
The reactions of iodine monoxide (IO) with sulfur-containing compounds, which are important for the atmospheric chemistry, are studied. An attempt is made to distinguish between the heterogeneous and homogeneous reaction pathways. It is shown that, under the experimental conditions, the reactions proceed on the wall and generate iodine atoms into the gas phase. It is found that, at room temperature, the rate constants for the gas-phase reactions of IO with (CH3)2S and H2S are lower than 2.5 × 10−14 and 8.0 × 10−14 cm3 molecule−1 s−1, respectively; the rate constant for the gas-phase reaction of iodine monoxide with SO2 ≤ 5.6 × 10−15 cm3 molecule−1 s−1.  相似文献   

2.
The rate constant of the reaction between iodomethane and chlorine atoms at 323 K, measured by the resonance florescence method under jet stream conditions as the iodine atom yield, is k 1I = (2.9±0.6) × 10−12 cm3 molecule−1 s−1. It is demonstrated experimentally that this reaction takes place mainly on the reactor wall.  相似文献   

3.
The reaction of iodine monoxide with chlorine monoxide resulting in atom escape to the gas phase is studied at T = (303 ± 5) K and P = 2.5 Torr using a flow setup for measuring the resonance fluorescence signals of atomic iodine and chlorine. The heterogeneous reaction between chlorine monoxide and iodine monoxide occurring at the reactor surface covered with an F32-L Teflon-like compound and treated by the reaction products is characterized by the rate constant k = (4.9 ± 0.2) × 10–11 cm3 molecule–1 s–1. This value is substantially higher than the rate constant for the homogeneous reaction IO· + ClO· (k 1 1 × 10–12 cm3 molecule–1 s–1).  相似文献   

4.
The rate constants of the reactions of the chlorine atom with C3F7I (k 1) and CF3I (k 2) have been measured using the resonance fluorescence of chlorine atoms in a flow reactor at 295 K: k 1 = (5.2 ± 0.3) × 10−12 cm3 molecule−1 s−1 and k 2 = (7.4 ± 0.6) × 10−13 cm3 molecule−1 s−1. No iodine atoms have been detected in the reaction products.  相似文献   

5.
The effect of several solvents on the selectivity of the free radical chlorination of 1,1-dichloroethane and 1-chloropropane is studied. The selective action of aromatic solvents on free radical chlorination is explained. This explanation implies that the process involves solvated chlorine atoms and their donor-acceptor complexes with aromatic molecules (ArH→Cl) as intermediates. Using the findings of this work and previous studies, the ratios of the rate constants for hydrogen-atom abstraction from different positions in chloroethane, 1,1-dichloroethane, 1-chloropropane, and 2-chloropropane by solvated chlorine atoms and ArH→-Cl complexes are determined. The differences in the activation energies of the competitive hydrogenatom abstractions from different positions in substrates by the ArH→Cl complexes and solvated chlorine radicals correlate with two HOMO energies of solvent and substrate molecules. The isokinetic relationship is found for all the systems under study (the isokinetic temperature, 523 K).  相似文献   

6.
Radiation chemical reactions ofOH, O•−, N3 and e aq t- witho- and m-hydroxycinnamic acids were studied. The second-orderrateconstantsforthereaction ofOH with ortho and meta isomers in buffer solution at pH7 are 3.9±0.2 × 109 and 4.4 ± 0.3 × 109 dm3 mol-1 s-1 respectively. At pH 3 the rate with the ortho isomer was halved (1.6 ± 0.4 × 109 dm3 mol-1 s-1) but it was unaffected in the case of meta isomer (k = 4.2±0.6 × 109dm3mol-1 s-1). The rate constant in the reaction of N3 with the ortho isomer is lower by an order of magnitude (k = 4.9 ± 0.4 × 108 dm3 mol-1s-1). The rates of the reaction of e aq t- with ortho and meta isomers were found to be diffusion controlled. The transient absorption spectrum measured in theOH witho-hydroxycinnamic acid exhibited an absorption maximum at 360 nm and in meta isomer the spectrum was blue-shifted (330 nm) with a shoulder at 390 nm. A peak at 420 nm was observed in the reaction of Obb−with theo-isomer whereas the meta isomer has a maximum at 390 and a broad shoulder at 450 nm. In the reaction of the absorption peaks were centred at 370–380 nm in both the isomers. The underlying reaction mechanism is discussed.  相似文献   

7.
In neutral aqueous solution of (phenylthio)acetic acid, hydroxyl radical is observed to react with a bimolecular rate constant of 7.2 × 10-1 dm3 mols and the transient absorption bands are assigned toOH radical addition to benzene and sulphur with a rough estimated values of 50 and 40% respectively. The reaction of theOH radical with diphenyl sulphide (k = 4.3 × 108 dm3 mol−1 s−1) is observed to take place with formation of solute radical cation, OH-adduct at sulphur and benzene with estimated values of about 12, 28 and 60% respectively. The transient absorption bands observed on reaction ofOH radical, in neutral aqueous solution of 4-(methylthio)phenyl acetic acid, are assigned to solute radical cation (λmax = 550 and 730 nm), OH-adduct at sulphur (λmax = 360 nm) and addition at benzene ring (λmax = 320 nm). The fraction ofOH radical reacting to form solute radical cation is observed to depend on the electron-withdrawing power of substituted group. In acidic solutions, depending on the concentration of acid and electron-withdrawing power, solute radical cation is the only transient species formed on reaction ofOH radical with the sulphides studied.  相似文献   

8.
The morphological transformation process of gold nanorods (Au-NRs) resulting from the reaction between tetracycline and iodine was monitored by the plasmon resonance absorption (PRA) spectra and the scanning electron microscope (SEM) images. It was found that iodine could fuse Au-NRs into sphericity with the lower aspect ratio and blue shift of the longitudinal PRA band. It was found, however, that the presence of tetracycline, since it can react with I2, decreases the effective concentration of I2 and its fusion effect on Au-NRs. As a result, the longitudinal PRA of Au-NRs shifts to longer wavelength linearly with increasing the concentration of tetracycline. With that, tetracycline can be detected in the range of 5.0×10∮5- 5.0×10−4 mol·L−1, with a limit of determination (LOD) of 2.4×10−6 mol·L−1 (3σ). Most foreign substances in the samples did not interfere in the detection, and tetracycline in the synthetic samples could be detected with the recovery in the range of 92.8%–107.2%, and RSD lower than 4.3%. The concentration of tetracycline in milk detected with standard addition method was so low that it accorded with the safety regulation. Supported by the National Natural Science Foundation of China (Grant Nos. 30570465 and 20425517)  相似文献   

9.
The appearance of oscillations depends critically on the pH for a closed system of ClO2–I2–ethyl acetoacetate in the absence of sulfuric acid, and was investigated by determining the absorbance of I3 with reaction time at 280 nm. The pH should be 2.2–3.8. The initial concentration of ethyl acetoacetate, chlorine dioxide, iodine, and sulfuric acid has great influence on the oscillation at 581 nm for I3–starch complex (SI3). The oscillation occurs as long as the reactants are mixed at 280 nm. There is no pre-oscillatory period. However, at 581 nm, there is an induction period. The curve’s shape at 581 nm is very different from that at 280 nm. The oscillation becomes more obvious by adding starch at 581 nm for I3–starch complex (SI3) than that observed without adding starch at 280 nm. The oscillation curve is more regular and smooth by adding starch at 581 nm than that without adding starch at 280 nm. The amplitude and the number of oscillations are associated with the initial concentration of reactants. The higher the initial concentration of ethyl acetoacetate, the bigger the amplitude. Also, the number of oscillations becomes small. An opposite influence exists for chlorine dioxide and iodine. The higher the initial concentration of sulfuric acid, the bigger the amplitude. Also, the number of oscillations becomes large. The equations for the triiodide ion reaction rate changing with reaction time and the initial concentrations on the oscillation stage were obtained. The intermediates were detected by the online FTIR analysis. Based upon the experimental data in this work and in the literature, a plausible reaction mechanism was proposed for the oscillation reaction.  相似文献   

10.
Summary Pulsed laser photolysis coupled with time-resolved UV-absorption monitoring of CH3COradicals was applied to obtain the rate constant, k1, for the reaction CH3CO+ HBr → CH3C(O)H + Br (1); k1(298 K) = (3.59 ± 0.23 (2σ))x10-12cm3molecule-1s-1. Utilization of k1in a third law procedure has provided the standard enthalpy of formation value ofDfH°298(CH3CO) = -10.04 ± 1.10 (2σ) kJ mol-1in excellent agreement with a very recent IUPAC recommendation.  相似文献   

11.
It is shown experimentally that Cl appreciably accelerates ozone decomposition in water (τ1/2 = 1.5 h versus 6 h in pure water). The decomposition of ozone in NaCl solutions includes the reversible reaction of ozone with the chloride ion (O3 + Cl → O3 + Cl) as the key step, which is followed by the development of a chain reaction in which chain propagation is performed alternately by the chlorine atom Cl and its monoxide ClO. The current concentrations of the chlorine atom are rather low ([Cl] ∼ 10−14 mol/l). The overall process is satisfactorily described by a first-order rate law with respect to ozone. The decomposition of ozone in aqueous solutions of NaCl is not accompanied by the formation of products other than oxygen. In particular, no noticeable amounts of hypochlorites and chlorates are observed. This is particularly significant for medicinal applications of ozonized isotonic solutions.  相似文献   

12.
The molecular geometry and electronic structure of hydroxy-substituted naphthazarin (NZ)-7-ethyl-2,3,5,6,8-pentahydroxy-1,4-naphthoquinone (echinochrome A, (Et)NZ(β-OH)3, 1) were calculated by the B3LYP/6-311G(d) method. The influence of the (i) character of the β-OH groups dissociation and (ii) conformational mobility of molecule 1 and the anions, radicals, and radical anions derived from 1 on the energy of their reactions with hydroperoxyl radical was studied by the (U)B3LYP/6-31G and (U)B3LYP/6-311G(d) methods. The enol-enolic tautomerism due to the transfer of hydrogen atoms of α-OH groups and rotational isomerism of the β-OH groups at the C(2) and C(3) atoms and of the α-OH groups at the C(5) and C(8) atoms were studied. The equilibrium in the gas-phase reaction 1 + OOH ⇄ (Et)(HO-β)2NZ(β-O) + HOOH (1) (quenching of hydroperoxyl radical) is shifted to the separated reagents. Heterolysis of the O—H bond in one of the three β-hydroxy groups considerably reduces the energy of subsequent O—H bond homolysis in either of the two remaining β-hydroxy groups. As a consequence, the reaction (Et)(HO-β)2NZ(β-O) + OOH ⇄ (Et)(HO-β,O-β)NZ(β-O) + HOOH (2) (quenching of hydroperoxyl radical) becomes exothermic and the equilibrium is shifted to the formation of hydrogen peroxide. The Gibbs energy gain in reaction (2) varies from −6.4 to −10.9 kcal mol−1 depending on which β-hydroxy group is involved in the O—H bond homolysis. Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 3, pp. 400–415, March, 2007.  相似文献   

13.
The mechanisms of reactions between CC13OO radical and quercetin, rutin and epigallocatechin gallate (EGCG) have been studied using pulse radiolytic technique. It is suggested that the electron transfer reaction is the main reaction between CC13OO radical and rutin, EGCG, but there are two main pathways for the reaction of CC13OO radical with quercetin, one is the electron transfer reaction, the other is addition reaction. The reaction rate constants were determined. It is proved that quercetin and rutin are better CC13OO radical scavengers than EGCG.  相似文献   

14.
Pulse radiolysis of 2-Mercaptobenzothiazole (2-MBT) has been undertaken in aqueous solution. The semi-oxidized species formed at pH 4.5 due to the reaction of OH, Br2 •− and N3 and at pH 10.5 with OH yielded a spectrum with λmax = 348 and 595 nm. These semi-oxidized species were able to oxidize phenothiazine drugs (Eo⋟0.8 V). Reducing species such as eaq , CO2 •− and H atoms react with 2-MBT resulting in the formation of a transient having λmax = 350 nm and reducing in nature. Kinetic and spectroscopic data of interest are reported.  相似文献   

15.
The recombination of carbon monoxide and oxygen atoms was studied in reflected shock waves in H2:O2:CO:Ar = 0.1:1:24:75 with 1300 < T5 2200 K and 2 < P5 < 4 atm. Reaction progress was monitored by observations of the carbon monoxide flame spectrum near 435 nm and carbon dioxide thermal emission near 4.2 μm. Data analysis was accomplished with the aid of computer modeling using a 27-reaction mechanism. Computer modeling experiments also showed that these measurements were sensitive primarily to the rate of the reaction CO + O + M = CO2 + M and only slightly sensitive to the rates of other reactions. The best fit to the data was achieved with a rate constant for this reaction of 7.7 × 10?35 exp[19 kJ/RT] cm6 s for the temperature range of these experiments. Correlation of this result and previous data covering the temperature range 250 < T < 11,000 K confirms that this recombination reaction is governed by a nonadiabatic curve crossing with an activation barrier of about 20 kJ and subsequent deactivation of a singlet CO2 molecule.  相似文献   

16.
Laser flash photolysis (at 248 or 308 nm) or aryl iodides in water or water/methanol solutions produces iodine atoms and phenyl radicals. Iodine atoms react rapidly with added I? to form I2? but do not react rapidly with O2 (k ? 107 L mol?1 s?1). Iodine atoms oxidize phenols to phenoxyl radicals, with rate constants that vary from 1.6 × 107 L mol?1 s?1 for phenol to about 6 × 109 L mol?1 s?1 for 4-methoxyphenol and hydroquinone. Ascorbate and a Vitamin E analogue are also oxidized very rapidly. N-Methylindole is oxidized by I atoms to its radical cation with a diffusion-controlled rate constant, 1.9 × 1010 L mol?1 s?1. Iodine atoms also oxidize sulfite and ferrocyanide ions rapidly but do not add to double bonds. The phenyl radicals, produced along with the I atoms, react with O2 to give phenylperoxyl radicals, which react with phenols much more slowly than I atoms. © 1995 John Wiley & Sons, Inc.  相似文献   

17.
The following reaction rate constants of oxygen atoms with iodomethane and chlorine were measured using resonance fluorescence under jet conditions at 298 K: k 1 = (2.4 ± 0.5) × 10–15 and k 2 = (6.9 ± 0.2) × 10−14 cm3/s, respectively.  相似文献   

18.
All gas-phase iodine laser (AGIL) powered by the decomposition of nitrogen trichloride (NCl3) is studied. This reaction scheme uses commonly available reagents and reaction paths are milder than the previously studied azide-based AGIL. Theoretical studies revealed the necessary operational conditions for achieving positive gain. An apparatus is made based on the results of the theoretical works. Positive gain at iodine I(2 P 1/2)-I(2 P 3/2) transition is observed for the first time. The article is published in the original.  相似文献   

19.
The experimental activation energies of the R + O = CR1R2 and RO + CH2 = CHR1 addition reactions are analyzed within the framework of the parabolic model of the bimolecular addition reaction. The activation energy also depends on the dissociation energy of the forming C-O bond and on the reaction enthalpy: the higher the dissociation energy, the higher the activation energy. The empirical relationshipr e J..D e = 0.97 x 10-13 m kJ.-1 mol is found for H, Cl, Br and RO radical addition to multiple C=C and C=O bonds (re is the distance between the peaks of the intersecting parabolic curves). This is due to the effect of the triplet repulsion on radical addition. The interaction of polar groups and the steric effect also influence the activation energy.  相似文献   

20.
The OH and the NO2 radicals generated pulse radiolytically in N2O-saturated aqueous solution at pH 8–8.5 oxidize Mesna to form the corresponding thiyl radicals which on reaction with thiolate ions form an RSSR type of transient with λmax = 420 nm. The rate constants for the formation of these transients were determined. In the absence of O2 at pH=6, the RS radicals formed show an absorption maximum at 360 nm and an ε=200±50 dm3 mol−1 cm−1. The rate constant k (OH+RSH) was 6×109 dm3 mol−1 s−1 as determined from competition kinetics. In the presence of O2 the Mesna thiyl radical was seen to rapidly add oxygen to form an RSOO type of species with λmax = 535 nm, ε=700±50 dm3 mol−1 cm−1 and k (RS+O2)=1.3×108 dm3 mol−1 s−1. Both the RS and the RSOO radicals formed by the oxidation of Mesna were able to abstract H-atoms from ascorbate ions and k(RS +AH)=~k(RSOO+AH)=~6−7×108 dm3 mol−1 s−1-. Moderately strong oxidants like CCl3OO and the (CH3)3CO radicals, having a reduction potential of +1.4−1.6 V vs NHE were unable to oxidize Mesna. The results thus reflect on the pro- and anti-oxidant properties of Mesna.  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