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.
Interatomic distances in the transition state were estimated for the reactions of radical abstraction: H· + H2, H· + HCl, H· + CH4, N·H2 + NH3, HO· + H2O, HO2 · + HOOH, and C·H3 + SiH4. The calculation was performed by the quantum-chemical density functional method or coupled clusters method (QCH), as well as by the methods of intersecting parabolas (IPM) and Morse curves (IMM), using experimental data (activation energies and reaction enthalpies). The results of the latter two methods are close to the quantum-chemical calculation and differ only by the increment a: r(IPM or IMM) = a + r(QCH), where a = –4.5·10–12 m for IPM and a = +1.9·10–12 m for IMM.  相似文献   

2.
Indirect electrooxidation of phenol, formaldehyde, and maleic acid in cells with and without a cation-exchange membrane, with a platinum anode and a gas-diffusion carbon black cathode, which generates hydrogen peroxide from molecular oxygen, proceeds with high efficiency and various oxidation depths, which depend on the intermediate nature: the process involving HO 2 - occurs selectively and yields target products, while the formation of HO2 · and HO· leads to the destruction of organic compounds to CO2 and H2O.  相似文献   

3.
Summary The kinetics of reaction of HO with [Ru(TPT)2]2+ and [Fe(TPT)2]2+ have been studied in detail. The former participates in an equilibrium with HO yielding a pseudo-base by attack at the ligand and, at very high concentrations of HO, dissociates to yield pure TPT quantitatively. [Fe(TPT)2]2+ dissociates rapidly in basic solution, even at 273 K, however, [Fe(TPT)(TPT · OH)]+ does in fact exist and the Fell and Rull reactions are quite similar, although that of Fell is much faster. The implications of these findings for the dissociation of [Fe(TPT)2]2+ over a wide range of pH are discussed.Patt XX, ref. 1.  相似文献   

4.
The concepts of the selective inhibition of reactions with the participation of HO 2· radicals by nitrobenzene in accordance with a cyclic mechanism were supported using cyclohexanol oxidation as an example. It was demonstrated that the nitroxyl radical generated from nitrobenzene selectively reacts with HO2 · to form hydrogen peroxide and nitrobenzene. A decrease in the rates of oxidation of esters, carboxylic acids, and ketones with specially chosen structures in the presence of nitrobenzene, as well as the detection of H2O2 and corresponding ,-unsaturated compounds among the reaction products, indicated that the degradation of -peroxyl radicals of the above compounds occurred under conditions of the liquid-phase oxidation of organic substances that result in the formation of the HO2· radical and an unsaturated compound.Translated from Kinetika i Kataliz, Vol. 45, No. 6, 2004, pp. 814–820.Original Russian Text Copyright © 2004 by Nepomnyashchikh, Nosacheva, Perkel.  相似文献   

5.
Aqueous solutions of nitrilotriacetic acid (NTA) were irradiated with gamma-rays. In deaerated acidic solutions G (IDA, iminodiacetic acid) was found to be 3.0 and in aerated solutions 2.7. Both H and OH radicals abstracted alpha hydrogen from the NZA molecule. The dehydrogenated radical disproportionated to NTA and IDA; however in presence of air, the radical added with O2 to give peroxy intermediate which was hydrolyzed to IDA and HO2. The rate constants, for the reaction of OH-radical with NTA at pH 2.0, 6.0 and 10.0 as determined by competition kinetic methods were 0.61·108, 5.5·108 and 42·108 dm3·mol–1·s–1, respectively. These indicated that the unprotonated form of NTA is more reactive than its protonated form. This has been attributed to the deactivation of alpha-hydrogen centers by protons through inductive effect.  相似文献   

6.
A discharge-flow apparatus with resonance fluorescence and chemiluminescence detection has been used to monitor O2(b 1σ) production from several reactions of the HO2 radical at 300 K and 1-torr total pressure. O2(b), HO2, and OH were observed when F atoms were added to H2O2 in the gas phase. Signal strengths of O2(b) were proportional to initial concentrations of H2O2 and HO2. These observations were analyzed by using a simple three step mechanism and a more complete computer simulation with 22 reaction steps. The results indicate that the F + HO2 reaction yields O2(b) with an efficiency of (3.6 ± 1.4) × 10?3. By monitoring [O2(b)] and [HO2] upon addition of an excess second reactant to HO2, O2(b) yields from the reactions of HO2 with O, Cl, D, H, and OH were found to be <1 × 10?2, <5 × 10?4, <2 × 10?3, <8 × 10?3, and <1 × 10?3, respectively. Yields of O2(b) from the HO2 ± HO2 reaction were found to be less than 3 × 10?2.  相似文献   

7.
The possible reactions of HO2 with five ketones were studied using a flow tube reactor equipped with a laser magnetic resonance detector. We did not observe reactive loss of HO2 in any of the five reactions. We place upper limits of <8 × 10−16, <7 × 10−16, <5 × 10−16, <4 × 10−16, and <9 × 10−16 (in units of cm3; molecule−1 S−1) at 298 K for the reactions of HO2 with CH3COCH3, CH3COC2H5, CH3COC3H7, C2H5COC2H5, and CH3COC4H9, respectively, to give products other than an adduct. We conclude that their reactions with HO2 are unlikely to be important loss processes for ketones in the atmosphere. © 2000 John Wiley & Sons, Inc. Int J Chem Kinet 32: 573–580, 2000  相似文献   

8.
Enthalpies of the addition of radicals HO, HO2 , CH3 , CH3O, CH3O2 to the C=0 bond of model aldehydes, ketones, acids and esters have been calculated by thermodynamic and quantum chemical methods. It has been shown that the exothermic effect of these reactions decreases in the row HO>CH3O>CH3 >HOO>CH3OO.  相似文献   

9.
The process of phenol oxidation on a boron-doped diamond electrode (BDD) is studied in acidic electrolytes under different conditions of generation of active oxygen forms (AOFs). The scheme of phenol oxidation known from the literature for other electrode materials is confirmed. Phenol is oxidized through a number of intermediates (benzoquinone, carboxylic acids) to carbon dioxide and water. Comparative analysis of phenol oxidation rate constants is performed as dependent on the electrolysis conditions: direct anodic oxidation, with oxygen bubbling, and addition of H2O2. A scheme is confirmed according to which active radicals (OH·, HO2·, HO2) are formed on a BDD anode that can oxidize the substrate which leads to formation of organic radicals interacting with each other and forming condensation products. Processes with participation of free radicals (chain-radical mechanism) play an important role in electrochemical oxidation on BDD. Intermediates and polymeric substances (polyphenols, quinone structures, and resins) are formed. An excess of the oxidant (H2O2) promotes a more effective oxidation of organic radicals and accordingly inhibition of the condensation process.  相似文献   

10.
Atmospheric pressure rate coefficients for the loss of HO2, CH3O2, and C2H5O2 radicals to the wall of a ¼″ Teflon tube have been measured. In dry air, they are 2.8 ± 0.2 s−1 for HO2 and 0.8 ± 0.1 s−1 for both CH3O2 and C2H5O2 radicals. The rate coefficient for HO2 loss increases markedly with the relative humidity of the air; however, the organic radicals show no such dependence. These data are used in a kinetic model of the radical amplifier chemistry to investigate the reported sensitivity to water concentration. The increased wall loss accounts for only some of the observed water dependence, suggesting there is an unreported water contribution to the gas phase chemistry. Including the reaction of the HO2/water adduct with NO to yield HNO3 or HOONO into the mechanism is shown to provide a better simulation of the observed water dependence of the radical detector. This reaction would also be important in atmospheric chemistry as it provides an additional loss mechanism for both radicals and NOx. ©1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 145–152, 1999  相似文献   

11.
The kinetics of the self-reactions of HO2, CF3CFHO2, and CF3O2 radicals and the cross reactions of HO2 with FO2, HO2 with CF3CFHO2, and HO2 with CF3O2 radicals, were studied by pulse radiolysis combined with time resolved UV absorption spectroscopy at 295 K. The rate constants for these reactions were obtained by computer simulation of absorption transients monitored at 220, 230, and 240 nm. The following rate constants were obtained at 295 K and 1000 mbar total pressure of SF6 (unit: 10−12 cm3 molecule−1 s−1): k(HO2+HO2)=3.5±1.0, k(CF3CFHO2+CF3CFHO2)=3.5±0.8, k(CF3O2+CF3O2)=2.25±0.30, k(HO2+FO2)=9±4, k(CF3CFHO2+HO2)=5.0±1.5, and k(CF3O2+HO2)=4.0±2.0. In addition, the decomposition rate of CF3CFHO radicals was estimated to be (0.2–2)×103 s−1 in 1000 mbar of SF6. Results are discussed in the context of the atmospheric chemistry of hydrofluorocarbons. © 1997 John Wiley & Sons, Inc.  相似文献   

12.
An array of 2D isoreticular layers, viz. [Zn(atrz)X] (1·X; X=Cl, Br, I; atrz=3-amino-1,2,4-triazole anion), [Zn4(atrz)4(SCN)4·H2O] (1·SCN·H2O) and [Zn(trz)X] (2·X; X=Cl, Br, I; trz=1,2,4-triazole anion), have been hydrothermally synthesized and structurally characterized. Compounds 1·X and 1·SCN·H2O are constructed from binuclear planar Zn2(atrz)2 subunits and exhibit (4,4) topological network when the subunits are simplified as four-connected nodes. Based on changing the terminal counteranions X (X=Cl, Br, I, SCN), the average interlayer separations of 1·X and 1·SCN·H2O are enlarged, which equal to 5.851, 6.153, 6.651 and 8.292 Å, respectively. As a result, H2O molecules reside in the spaces between two adjacent layers of 1·SCN·H2O. 2 and 1 are the isomorphous structures. In common with 1, the interlayer separations of 2·X are widened with increasing the ion radius. Solid-state luminescence properties and thermogravimetric analyses of 1 and 2 were investigated, respectively.  相似文献   

13.
The reaction of OH? with O3 eventually leads to the formation of .OH radicals. In the original mechanistic concept (J. Staehelin, J. Hoigné, Environ. Sci. Technol. 1982 , 16, 676–681), it was suggested that the first step occurred by O transfer: OH?+O3→HO2?+O2 and that .OH was generated in the subsequent reaction(s) of HO2? with O3 (the peroxone process). This mechanistic concept has now been revised on the basis of thermokinetic and quantum chemical calculations. A one‐step O transfer such as that mentioned above would require the release of O2 in its excited singlet state (1O2, O2(1Δg)); this state lies 95.5 kJ mol?1 above the triplet ground state (3O2, O2(3Σg?)). The low experimental rate constant of 70 M ?1 s?1 is not incompatible with such a reaction. However, according to our calculations, the reaction of OH? with O3 to form an adduct (OH?+O3→HO4?; ΔG=3.5 kJ mol?1) is a much better candidate for the rate‐determining step as compared with the significantly more endergonic O transfer (ΔG=26.7 kJ mol?1). Hence, we favor this reaction; all the more so as numerous precedents of similar ozone adduct formation are known in the literature. Three potential decay routes of the adduct HO4? have been probed: HO4?→HO2?+1O2 is spin allowed, but markedly endergonic (ΔG=23.2 kJ mol?1). HO4?→HO2?+3O2 is spin forbidden (ΔG=?73.3 kJ mol?1). The decay into radicals, HO4?→HO2.+O2.?, is spin allowed and less endergonic (ΔG=14.8 kJ mol?1) than HO4?→HO2?+1O2. It is thus HO4?→HO2.+O2.? by which HO4? decays. It is noted that a large contribution of the reverse of this reaction, HO2.+O2.?→HO4?, followed by HO4?→HO2?+3O2, now explains why the measured rate of the bimolecular decay of HO2. and O2.? into HO2?+O2 (k=1×108 M ?1 s?1) is below diffusion controlled. Because k for the process HO4?→HO2.+O2.? is much larger than k for the reverse of OH?+O3→HO4?, the forward reaction OH?+O3→HO4? is practically irreversible.  相似文献   

14.
The rate constants for the addition of ·CH(Ph)CH2CCl3, ·CH2Ph, ·CH2Prn, and ·CCl3 radicals to the ethyl 2-cyanoacrylate molecule were determined by ESR spectroscopy using the spin trapping technique.  相似文献   

15.
The kinetics of the title reactions were investigated in a discharge flow tube by using laser magnetic resonance detection of HO2. The upper limits for the bimolecular rate constants for the reactions of HO2 with H2S (k1), CH3SH (k2), and CH3SCH3 (k3) are <3 × 10?15, <4 × 10?15, and <5 × 10?15 cm3 molecule?1 s?1, respectively, at 298 K. Our upper limit for k1 is three orders of magnitude lower than the previously reported value. Measurements at higher temperatures also yield similar upper limits. Our results suggest that HO2 is not an important oxidant for these reduced compounds in the atmosphere. © 1994 John Wiley & Sons, Inc.  相似文献   

16.
A new tetradentate imidazolate ligand 1,1′,1″,1′′′-(2,2′,4,4′,6,6′-hexamethylbiphenyl-3,3′,5,5′-tetrayl)tetrakis(methylene)(1H-imidazole) (L) and four Ag(I)/Cu(I) coordination polymers, namely [(MCN)3L]n (1: M=Ag; 2: M=Cu), and [(MSCN)2L]n (3: M=Ag; 4: M=Cu) are described. All four new coordination polymers were fully characterized by infrared spectroscopy, elemental analysis and single-crystal X-ray diffraction. Compound 1 features a 3D supramolecular framework constructed by 1D chains through inter-chain Ag-N(CN) and inter-layer Ag-N(L) weak interactions with an uninodal 66 topology. Complex 2 presents a 3D framework characterized by a tetranodal (3,4)-connected (3·4·5·102·11)(3·4·5·6·7·9)(3·6·7)(6·102) topology. Complexes 3 and 4 are isostructural, and both have a 3D network of trinodal 4-connected (4·85)2(42·82·102)(42·84)2 topology. The luminescent properties for these compounds in the solid state as well as the possible ferroelectric behavior of 1 are discussed.  相似文献   

17.
It is demonstrated by simultaneous analysis of experimental data and the results of numerical simulation that, at 100 kPa, the H2/O2 mixture self-ignites only owing to the occurrence of a chain avalanche. Self-heating becomes significant in the course of chain combustion and strengthens the chain avalanche. The accelerating decomposition of H2O2, which results from the reaction between HO2 and H2, contributes to chain branching and determines the character of the chain thermal explosion. The role of the HO2 radical depends on the chain process conditions, consisting of chain termination as a result of the partial loss of HO2 at the initial self-acceleration stage, H atom regeneration, and participation in the second branched cycle under conditions of accelerated H2O2 decomposition.  相似文献   

18.
Electron pulse radiolysis at ?298°K of 2 atm H2 containing 5 torr O2 produces HO2 free radical whose disappearance by reaction (1), HO2 + HO2 →H2O2 + O2, is monitored by kinetic spectrophotometry at 230.5 nm. Using a literature value for the HO2 absorption cross section, the values k1 = 2.5×10?12 cm3/molec·sec, which is in reasonable agreement with two earlier studies, and G(H) G(HO2) ?13 are obtained. In the presence of small amounts of added H2O or NH3, the observed second-order decay rate of the HO2 signal is found to increase by up to a factor of ?2.5. A proposed kinetic model quantitatively explains these data in terms of the formation of previously unpostulated 1:1 complexes, HO2 + H2O ? HO2·H2O (4a) and HO2 + NH3? HO2·NH3 (4b), which are more reactive than uncomplexed HO2 toward a second uncomplexed HO2 radical. The following equilibrium constants, which agree with independent theoretical calculations on these complexes, are derived from the data: 2×10?20?K4a?6.3 × 10?19 cm3/molec at 295°K and K4b = 3.4 × 10?18 cm3/molec at 298°K. Several deuterium isotope effects are also reported, including kH/kD = 2.8 for reaction (1). The atmospheric significance of these results is pointed out.  相似文献   

19.
The microphase adsorption-spectral correction (MPASC) technique is described and applied to the study of the interactions of Evans blue (EB) with cetyltrimethylammonium bromide (CTAB) and with four proteins: bovine serum albumin (BSA), myoglobin (Mb), hemoglobin (Hb) and ovalbumin (OVA). EB can be adsorbed on a cationic surfactant and on protein by electrostatic force and the aggregation obeys the Langmuir isotherm. Results have shown that the products are formed as follows: monomer aggregate EB·CTAB, micellar aggregate (EB·CTAB)78 and protein aggregates (EB68·BSA), (EB14·OVA), (EB126·Mb) and (EB58·Hb). The adsorption constant of the aggregates are calculated to be KEB·CTAB=2.95×106, KEB68·BSA=3.40×104, KEB14·OVA=5.20×102, KEB126·Mb=6.81×102 and KEB58·Hb=5.73×102, respectively. The aggregation of EB in proteins is sensitive in the presence of CTAB and selective in the presence of EDTA and it has been applied to the analysis of samples with satisfactory results.  相似文献   

20.
Layered barium phosphonate, synthesized by combining the metallic salt with a phenylphosphonic acid solution, yielded Ba(HO3PC6H5)2 ·H2O (BaPP), which gives the corresponding anhydrous compound on heating. n-Alkylmonoamines intercalation into the crystalline lamellar precursor resulted in compounds having the general formula Ba(HO3PC6H5)2 ·xH2N(CH2) n CH3 ·(1−x)H2O (n=1–5). The intense infrared bands in the 1160–695 cm−1 interval confirmed the presence of the phosphonate groups attached to the inorganic layer, with sharp and intense peaks in X-ray diffraction patterns for both hydrated and anhydrous compounds. The thermogravimetric curves for both supports showed the release of water molecules and the organic moiety in distinct stages to yield a final Ba(PO3)2residue. An additional amine mass loss steps was observed for the corresponding aminated compounds. One isolated DSC peak found in the layered precursor compound contrasts by its absence in the anhydrous form and the 3P NMR spectrum presented one peak for attached phenylphosphonate groups centered at 12.4 ppm. An increase in carbon and hydrogen percentages for intercalated compounds followed the amine size chain with a corresponding decrease in nitrogen percentage. The interlayer distance (d) correlates linearly with the number of carbon atoms (n c ) of the alkylamine chains, d=1467 + 62n c and d=1688 + 60n c , for the hydrated and anhydrous compounds, respectively, permitting inference of the interlayer distance for an unknown amine.This revised version was published online in July 2005 with a corrected issue number.  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