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 140 毫秒
1.
The mechanism of the photolysis of formaldehyde was studied in experiments at 3130 Å and in the pressure range of 1–12 torr at 25°C. The experiments were designed to establish the quantum yields of the primary decomposition steps (1) and (2), CH2O + hν → H + HCO (1): CH2O + hν → H2 + CO (2), through the effects of added isobutene, trimethylsilane, and nitric oxide on ΦCO and Φ. The ratio ΦCO/Φ was found to be 1.01 ± 0.09(2σ) and (Φ + ΦCO)/2 = 1.10 ± 0.08 over the range of pressures and a 12-fold change in incident light intensity. Isobutene and nitric oxide additions reduced Φ to about the same limiting value, 0.32 ± 0.03 and 0.34 ± 0.04, respectively, but these added gases differed in their effects on ΦCO. With isobutene addition ΦCO/Φ reached a limiting value of 2.3; with NO addition ΦCO exceeded unity. The addition of small amounts of Me3SiH reduced Φ to 1.02 ± 0.08 and lowered ΦCO to 0.7. These findings were rationalized in terms of a mechanism in which the “nonscavengeable,” molecular hydrogen is formed in reaction (2) with ?2 = 0.32 ± 0.03, while the “free radical” hydrogen is formed in reaction (1) with ?1 = 0.68 ± 0.03. In the pure formaldehyde system these reactions are followed by (3)–(5): H + CH2O → H2 + HCO (3); 2HCO → CH2O + CO (4); 2HCO → H2 + 2CO (5). The data suggest k4/k5 ? 5.8. Isobutene reduced Φ by the reaction H + iso-C4H8 → C4H9 (20), and the results give k20/k3 ? 43 ± 4, in good agreement with the ratio of the reported values of the individual constants k3 and k20.  相似文献   

2.
A kinetic study has been made of the 3130-Å photolysis of CH2O (8 torr) in O2-containing mixtures (0.02–8 torr) and in the presence of added CO2 (0–300 torr) at 25°C. Quantum yields of formation of H2, CO, and CO2 and the loss of O2 were measured. Φ and ΦCO were much above unity. In an explanation of these unexpected results, a new H-atom-forming chain mechanism was postulated involving HO2 and HO addition to CH2O: CH2O + hν → H + HCO (1) H + CH2O → H2 + HCO (3) H + O2 + M → HO2 + M (6) HCO + O2 → HO2 + CO (8) HO2 + CH2O → (HO2CH2O) → HO + HCO2H (15) HO + CH2O → H2O + HCO? (16); HCO? → H + CO (19) HO + CH2O → H2O + HCO (17) and HO + CH2O → HCO2H + H (18). When the results are rationalized in terms of this mechanism, the data suggest k16 ? k17 and k16/k18 ? 0.5. The data require that a reassessment of the relative rates of reactions (7) and (8) be made, since in the previous work HCO2H formation was used as a monitor of the rate of reaction (7) HCO + O2 + M → HCOO2 + M (7). The present data from experiments at P = 8 torr and P = 1–4 torr give k7[M]/(k7[M] + k8) ≥ 0.049 ± 0.017. These data coupled with the k8 estimates of Washida and coworkers give k7 ≥ (4.4 ± 1.6) × 1011 l2/mol2·sec for M = CH2O. The reaction sequence proposed here is consistent with the observed deterimental effect of O2 addition on the laser-induced isotope enrichment in HDCO. In additional studies of CH2O-O2-isobutene mixtures it was found that Φ was equal to ?2 as estimated in O2-free CH2O-isobutene mixtures. These results suggest that the increase in CO (ν = 1) product observed with O2 addition in CH2O photolysis does not result from perturbations in the fragmentation pattern of the excited CH2O, but it is likely that it originates in the occurrence of the exothermic reaction HCO + O2 → HO2 + CO (ν = 1).  相似文献   

3.
Alcohols are known to promote the disproportionation of Cu(I)X species into nascent Cu(0) and Cu(II)X. Therefore, alcohols are expected to be excellent solvents that facilitate the single‐electron transfer mediated living radical polymerization (SET‐LRP) mediated by nascent Cu(0) species. This publication demonstrates the ultrafast SET‐LRP of methyl acrylate initiated with bis(2‐bromopropionyloxy)ethane and catalyzed by Cu(0)/Me6‐TREN in methanol, ethanol, 1‐propanol, and tert‐butanol and in their mixture with water at 25 °C. The structural analysis of the resulting polymers by a combination of 1H NMR and MALDI‐TOF MS demonstrates the synthesis of perfectly bifunctional α,ω‐dibromo poly(methyl acrylate)s by SET‐LRP in alcohols. Moreover, this work provides an expansion of the list of solvents available for SET‐LRP. © 2008 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 46: 2745–2754, 2008  相似文献   

4.
Experimental quantum yields of the photolysis of formaldehyde at lambda > 310 nm are combined with absolute and relative rate calculations for the molecular elimination H2CO --> H2 + CO (1), the bond fission H2CO --> H + HCO (2), and the intramolecular hydrogen abstraction H2CO --> H ... HCO --> H2 + CO (3) taking place in the electronic ground state. Temperature and pressure dependencies of the quantum yields are analyzed with the goal to achieve consistency between experiment and modeling. Two wavelength ranges with considerably different properties are considered: 340-360 nm, where channel 1 competes with collisional deactivation of excited molecules, and 310-340 nm, which is dominated by the competition between the formation of radical and molecular products. The close relation between photolysis and pyrolysis of formaldehyde, such as analyzed for the pyrolysis in the companion paper, is documented and an internally consistent treatment of the two reaction systems is provided. The quantum yields are modeled and represented in analytical form such that values outside the available experimental range can be predicted to some extent.  相似文献   

5.
The wavelength dependence of HCO(0,0,0) formation in the photodissociation of acetaldehyde was measured using narrow (0.1 nm) bandwidth laser excitation and time-resolved intracavity laser detection (TRILD). A sharp energetic onset at 320 ± 1 nm (89.3 ± 0.3 kcal) for HCO(0,0,0) formation was found. The maximum concentration of HCO(0,0,0) occurs between 100 and 250 μs after excitation depending on the wavelength of excitation  相似文献   

6.
A spectrophotometric method has been used for the determination of the dissociation constant of P-nitroaniline indicator (which has a charge type A+B0) in various THF-water solvent mixtures at 25°C. The dissociation constant is found to decrease with the addition of THF which indicates that the basicity of the medium is increased by the addition of the organic solvent. The relative basicity of the medium is compared with other aqueous-organic solvent mixtures and the trend of decreasing basicity is in the order THF-H2O> DMF-H2O > AN-H2O > HAC-H2O.  相似文献   

7.
The alkyl nitrites, C2H5ONO, n-C3H7ONO, n-C4H9ONO, and i-C4H9ONO were photolyzed at 23°C in the presence of 15NO at 366-nm incident radiation. The quantum yields of the corresponding isotopically-enriched alkyl nitrites were measured by mass spectrometry. The results indicated that only part of the absorption leads to photodecomposition. The remainder forms an electronically excited state which isotopically exchanges with 15NO. The indicated reactions of the electronically excited state RONO*, are where k3/k2 = 0.50 ± 0.10, 0.62 ± 0.20, 0.42 ± 0.06, and 0.24 ± 0.03 torr, and that k2a/k2 = 1.0, 1.0, 0.64 ± 0.04, and 0.56 ± 0.03, respectively, for C2H5ONO, n-C3H7ONO, n-C4H9ONO, and i-C4H9ONO.  相似文献   

8.
The development of Cu(0)/TREN/CuBr2‐catalyzed SET‐LRP of VC initiated with CHBr3 in DMSO at 25 °C is reported. The use of CuBr2 additive allows for the first LRP of low molecular weight VC (target DP = 100), as well as lower Cu powder loading levels, improved Ieff and control in the synthesis of higher molecular VC, targeted degree of polymerization = 350, 700, 1,000, 1,400. 1H NMR and HSQC confirm the bifunctionality of CHBr3 as an initiator and suggest that deleterious side‐reactions such as the formation of allylic chlorides occur primarily at the onset of the reaction. © 2009 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 47: 4130–4140, 2009  相似文献   

9.
The surface tensions of water samples of various degrees of purity were measured at 100°C using the Wilhelmy slide method. The purest water, obtained by constant redistillation and overflowing of the distillate surface assayed at 60.22 dyn/cm which may be compared with 58.74 dyn/cm (σ = 5.994 mg/mm) reported in the literature. Samples of distilled water of lesser purity gave values between σ = 59.48 dyn/cm and σ = 60.16 dyn/cm. It is concluded that the accepted value at 100°C for water should be increased by 2.5%.  相似文献   

10.
The ionization and solution enthalpies of 2,4-dinitrophenol were measured calorimetrically at 25°C in water—DMSO mixtures ranging from 0.1 to 0.8 mole fraction of DMSO.

The greater acidity of 2,4-dinitrophenol with respect to 2,5-dinitrophenol is explained on the basis of the observation that in the anion the π-withdrawing resonance effect of the para nitro group stabilizes the benzene ring while in the undissociated molecule the contrary is true.  相似文献   


11.
The ionization and solution enthalpies of 2,5-dinitrophenol were measured calorimetrically at 25°C in water—DMSO mixtures ranging from 0.1 to 0.8 mole fraction of DMSO.

The effect of the nitro group in the ortho position seems to prevail over that of the nitro group in the meta position.

The greater acidity of dinitrophenol with respect to the mononitro isomers is explained on the basis of the interference of the nitro group in the metaposition on the interactions between the nitro group in the ortho position and the hydroxyl group or phenolate oxygen.  相似文献   


12.
Values of the emf for the cell: Na-glass|Na2Succ(m), Hg2Succ|Hg, have been measured. These values of E have been used with Hückel's extended Debye and Hückel equation to obtain the mean ionic activity coefficient values for Na2Succ in aqueous solutions of several concentrations, m. The values of γ± thus obtained have been adjusted to Pitzer's equation, with b = 1.2 and α = 2.0 to find Pitzer's parameters and then the values for both the osmotic coefficients and the activity of water. Finally, the standard potential of the Hg2Succ|Hg electrode has been determined and from this value the solubility-product equilibrium constant for mercurous succinate has also been calculated.  相似文献   

13.
A series of measurements of integral n-hexane vapor sorption at 25°C and moderate activity, in polystyrene microspheres of varying radii, exhibits typical non-Fickian behavior: nonoverlap of curves of fractional uptake versus (square root of time)/radius. The data are examined in light of a sorption isotherm indicating hexane solubility in excess of that predicted by the Flory–Huggins equation, up to the hexane activity at which the glass transition apparently occurs. A transport analysis is developed based on the assumption that below the transition temperature Tg the rate of sorption is limited by the rate of polymer chain relaxations induced by the penetrant, which facilitate hexane entry into, and immobilization in, glassy microvoids.  相似文献   

14.
The photolysis wavelength dependence of the nitrate radical quantum yield for peroxyacetyl nitrate (CH(3)C(O)OONO(2), PAN) is investigated. The wavelength range used in this work is between 289 and 312 nm, which mimics the overlap of the solar flux available in the atmosphere and PAN's absorption cross section. We find the nitrate radical quantum yield from PAN photolysis to be essentially invariant; Phi(NO3)(PAN) = 0.30 +/- 0.07 (+/-2sigma) in this region. The excited states involved in PAN photolysis are also investigated using ab initio calculations. In addition to PAN, calculations on peroxy nitric acid (HOONO(2), PNA) are performed to examine general photochemical properties of the -OONO(2) chromophore. Equation of motion coupled cluster calculations (EOM-CCSD) are used to examine excited state energy gradients for the internal coordinates, oscillator strengths, and transition energies for the n --> pi* transitions responsible for the photolysis of both PNA and PAN. We find in both molecules, photodissociation of both O-O and O-N bonds occurs via excitation to predissociative electronic excited states and subsequent redistribution of that energy as opposed to directly dissociative excitations. Comparison and contrast between experimental and theoretical studies of HOONO(2) and PAN photochemistry from this and other work provide unique insight on the photochemistry of these species in the atmosphere.  相似文献   

15.
An acceleration effect and selective monomer addition during RAFT copolymerization of the oppositely‐charged ionic monomers in dilute aqueous solution at 25 °C are reported. The reaction is conducted using a non‐ionic water‐soluble polymer as a macromolecular chain transfer agent under visible light irradiation. A fast iterative polymerization can be induced, even in dilute solution, by the favorable ionic interactions and in situ self‐assembly of zwitterionic growing chains. Selelctive monomer addition is achieved in the statistical copolymerization due to the ion‐pairing of the oppositely‐charged monomers, such as precisely the same reaction rates at a 1:1 of monomer ratio, otherwise a faster reaction of the minor monomer component over the major one. These behaviors open up an avenue towards the rapid synthesis of sequence‐controlled zwitterionic polyelectrolytes that can satisfy the demands of emerging biological applications.

  相似文献   


16.
A mechanistic comparison of the ATRP and SET‐LRP is presented. Subsequently, simulation of kinetic experiments demonstrated that, in the heterolytic outer‐sphere single‐electron transfer process responsible for the SET‐LRP, the activation of the initiator and of the propagating dormant species is faster than of the homolytic inner‐sphere electron‐transfer process responsible for ATRP. In addition, simulation experiments suggested that in both polymerizations the rate of deactivation is similar. In SET‐LRP, the Cu(II)X2/L deactivator is created by the disproportionation of Cu(I)X/L inactive species, while in ATRP its concentration is mediated by the bimolecular termination. The combination of higher rate of activation with the creation of deactivator via disproportionation provides, via SET‐LRP, an ultrafast synthesis of polymers with very narrow molecular weight distribution at room temperature. SET‐LRP is mediated by a catalytic amount of Cu(0), and under suitable conditions, bimolecular termination is virtually absent. Kinetic and simulation experiments have also demonstrated that the amount of water available in commercial solvents and monomers is sufficient to induce the disproportionation of Cu(I)X/L into Cu(0) and Cu(II)X2/L and, subsequently, to change the polymerization mechanism from ATRP to SET‐LRP. © 2006 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 45: 1835–1847, 2007.  相似文献   

17.
KNO2 III below ?13°C is monoclinic, space group P21 or P21/m, with a0, b0, c0 = 4.677, 9.650, 6.395 Å, β = 93.8° at ?35°C. There is a further phase transformation between ?35°C and ?100°C to a new phase KNO2 VII, which is also monoclinic, space group P21 or P21/m: with a0, b0, c0 = 8.397, 4.773, 7.644 Å, β = 112° at ?100°C. Both these phases appear to be ordered.  相似文献   

18.
19.
GAO  Qiangb 《中国化学》2009,27(7):1291-1294
基于氯过氧化物酶(CPO)的卤化活性分析,发现某些碱土金属(Ca2+, Mg2+)和过渡金属(Co2+, Ni2+)对CPO具有明显的激活及稳定化作用。例如25 ºC时与CPO在纯缓冲溶液中相比,在75 μmol·L-1 Ca2+,90 μmol·L-1 Mg2+,90 μmol·L-1 Ni2+及105 μmol·L-1 Co2+存在时CPO可分别获得1.33,1.37,1.34 及1.27倍的最大相对活性。而在55 ºC,没有金属离子存在时,CPO 30分钟后仅能保留40%的活性,但在Ca2+,Mg2+离子的介质中,CPO的活性可分别保留81% 和 75%。推测这是由于金属离子结合在CPO活性中心周围的酸-碱催化位点Glu183, His105 and Asp106上,通过底物浓集和诱导有利构象来激活CPO. 同时动力学研究表明金属离子对CPO的激活归因于催化效率(kcat)的提高,以及CPO对底物亲和性及选择性的改善。  相似文献   

20.
Formation of the chloro complexes of manganese(II), cobalt(II), nickel(II), copper(II) and zinc(II) in DMSO has been studied potentiometrically at 25°C. The con- centration stability constants for the ionic strength of 0.1 mol kg are derived and discussed. The stability of the divalent transition metal cations towards the chloride anion follows the sequence Mn &> Co &> Ni << Cu &> Zn disobeying the Irving-Williams series.  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