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 241 毫秒
1.
Carbon-13 hyperfine splittings equal to 41±3 gauss have been observed in the paramagnetic resonance of a mixture of C12H3 and C13H3 radicals produced by x-irradiation of CH3I at 77°k. The observed splitting provides strong evidence that CH3 is a planar molecule.  相似文献   

2.
ABSTRACT

We explored a novel doubly labelled water (DLW) method based on breath water (BW-DLW) in mice to determine whole body CO2 production and energy expenditure noninvasively. The BW-DLW method was compared to the DLW based on blood plasma. Mice (n?=?11, 43.5?±?4.6?g body mass (BM)) were administered orally a single bolus of doubly labelled water (1.2?g H218O kg BM?1 and 0.4?g 2H2O kg BM?1, 99 atom% (AP) 18O or 2H). To sample breath water, the mice were placed into a respiration vessel. The exhaled water vapour was condensed in a cold-trap. The isotope enrichments of breath water were compared with plasma samples. The 2H/1H and 18O/16O isotope ratios were measured by means of isotope ratio mass spectrometry. The CO2 production (RCO2) was calculated from the 2H and 18O enrichments in breath water and plasma over 5 days. The isotope enrichments of breath water vs. plasma were correlated (R2?=?0.89 for 2H and 0.95 for 18O) linearly. The RCO2 determined based on breath water and plasma was not different (113.2?±?12.7 vs. 111.4?±?11.0?mmol?d–1), respectively. In conclusion, the novel BW-DLW method is appropriate to obtain reliable estimates of RCO2 avoiding blood sampling.  相似文献   

3.
Relative kinetics of the reactions of OH radicals and Cl atoms with 3‐chloro‐2‐methyl‐1‐propene has been studied for the first time at 298 K and 1 atm by GC‐FID. Rate coefficients are found to be (in cm3 molecule?1 s?1): k1 (OH + CH2 = C(CH3)CH2Cl) = (3.23 ± 0.35) × 10?11, k2 (Cl + CH2 = C(CH3)CH2Cl) = (2.10 ± 0.78) × 10?10 with uncertainties representing ± 2σ. Product identification under atmospheric conditions was performed by solid phase microextraction/GC‐MS for OH reaction. Chloropropanone was identified as the main degradation product in accordance with the decomposition of the 1,2‐hydroxy alcoxy radical formed. Additionally, reactivity trends and atmospheric implications are discussed. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

4.
Hydrogen (H2) is known to be the fastest fuel to ignite among all practical combustion fuels. In this study, for the first time, longer ignition delay times (IDTs) for the H2 and H2 blended CH4 mixtures were measured compared to those for pure CH4. This work investigates the ignition characteristics of H2, CH4, and 50% CH4/50% H2 mixtures using a rapid compression machine at pressures ranging from 20 to 50 bar and at equivalence ratios (φ) from 0.5 to 2.0 in air in the temperature range 858–1080 K. The experimental IDTs are simulated using a newly updated kinetic mechanism, NUIGMech1.3, and good agreement is observed. At lower temperatures the IDTs of H2, CH4, and the 50% CH4/50% H2 mixtures are similar to one another, and the IDTs of the 50% CH4/50% H2 mixtures are longer than those for pure CH4 at temperatures below 930 K. At temperatures below 890–925 K, depending on the operating pressure and equivalence ratio, the hydrogen mixtures are the slowest to ignite, with IDTs being 2.5 times longer than those recorded for CH4 at a pressure of 40 bar at 890 K for φ = 1.0, and at 875 K for φ = 2.0. At low temperatures alkyl (Ṙ = ĊH3 and Ḣ) radicals add to O2 producing RȮ2 radicals, which then react with HȮ2 radicals forming ROOH (H2O2 and CH3OOH) and O2. For H2, the self-recombination of HȮ2 radicals leads to chain propagation which inhibits reactivity, whereas for CH4, the reaction between RȮ2 (CH3OȮ) and HȮ2 leads to chain branching, increasing reactivity. Furthermore, CH3OOH decomposes more easily to produce CH3Ȯ and ȮH radicals than does H2O2 to produce two ȮH radicals. Thus, mixtures containing higher H2 concentrations are slower to ignite compared to those with higher CH4 concentrations at low temperatures.  相似文献   

5.
Abstract

The electron spin resonance of γ-irradiated single crystals of methoxycarbonylcholine picrate hemihydrate, C7H16NO3 + · C6H2N3O7 ? · ½ H2O has been observed and analyzed for different orientations of the crystal in the magnetic field. The crystals have been investigated between 70 and 350 K. The spectra were found to be temperature independent and the radiation damage centers are attributed to – ?[Obar]OCH3 and –CH2CH2O? radicals. The g and hyperfine coupling constants were found to be almost isotropic with an average, g = 2.0060, a H1 = 4.4G for –CH2CH2O?, g = 2.0050, a H2 = 3.5G for –CH2CH2O? and g = 2.0045, a H = 3.5 G for –?[Obar]OCH3. These values indicate a long-range coupling between the unpaired electron and H protons.  相似文献   

6.
The anisotropic hyperfine coupling constants (AHCC) from the electron spin resonance (E.S.R.) spectra of a variety of atoms in organic radicals have been calculated by means of semiempirical molecular orbital wavefunctions in the INDO approximation. Hyperfine tensors involving 1H, 13C and 19F nuclei are obtained for the ?H, ?H3, CH3?H2, (CH3)3? hydrocarbon radicals, malonic acid radical, ?H2F, ?F2H, ?F3 and CF3?H2 radicals. The calculated values are compared with available experimental, non-empirical and semiempirical values for these radicals. All integrals of the operator entering the electronic contributions have been evaluated over Slater type orbitals. The introduction of deorthogonalized wavefunctions gives generally better calculated results. In particular, the tensor components of the 19F AHCC are in good agreement with the experimental results without the necessity of readjusting the effective nuclear charges.  相似文献   

7.
When an aqueous Au(III) solution containing 1-butanol was sonicated under Ar, Au(III) was reduced to Au(0) to form Au particles. This is because various reducing species are formed during sonication, but the reactivity of these species has not yet been evaluated in detail. Therefore, in this study, we analyzed the effects of Au(III) on the rates of the formation of gaseous and water-soluble compounds (CH4, C2H6, C2H4, C2H2, CO, CO2, H2, H2O2, and aldehydes), and the rate of Au(III) reduction as a function of 1-butanol concentration. The following facts were recognized: 1) for Au(III) reduction, the contribution of the radicals formed by the pyrolysis of 1-butanol was higher than that of the secondary radicals formed by the abstraction reactions of 1-butanol with ·OH, 2) ·CH3 and CO acted as reductants, 3) the contribution of ·H to Au(III) reduction was small in the presence of 1-butanol, 4) aldehydes and H2 did not act as reductants, and 5) the types of species that reduced Au(III) changed with 1-butanol concentration.  相似文献   

8.
Benzophenone ((C6H5)2CO) and decafluorobenzophenone ((C6F5)2CO) were applied to elucidate the photochemical reaction pathway of hydrogen peroxide (H2O2) with dimethylsulfoxide (DMSO). When a solution of benzophenone in DMSO was excited with the 355 nm laser light, three transient species were observed in the time-resolved electron paramagnetic resonance spectra: benzophenone ketyl (C6H5)2COH, methylCH3, and methylsulfinic methylCH2SOCH3 radicals. However, when decafluoro-benzophenone was used with DMSO, only ketyl and methylsulfinic methyl radicals were observed under the same experimental conditions. When the reaction of benzophenone and DMSO was carried out in the presence of H2O2, different time profiles ofCH3 radicals were observed. In the reaction of decafluorobenzophenone-DMSO-H2O2, the time profiles of the radicals were not affected by the presence of H2O2. Thus, these results verify thatCH3 radicals are regenerated in a cyclic pathway, in whichCH3 radicals attack H2O2. The regeneration pathway allows us to observe f-pair polarization throughout the lifetime ofCH3 radicals, which last several microseconds, an order of magnitude longer than theT 1 relaxation time ofCH3 radicals.  相似文献   

9.
Hydroperoxides and the corresponding peroxy radicals are important intermediates during the partial oxidation of methyl ethyl sulfide (CH3SCH2CH3) in both atmospheric chemistry and in combustion. Structural parameters, internal rotor potentials, bond dissociation energies, and thermochemical properties (ΔHfo, So and Cp(T)) of 3 corresponding hydroperoxides CH2(OOH)SCH2CH3, CH3SCH(OOH)CH3, CH3SCH2CH2OOH of methyl ethyl sulfides, and the radicals formed via loss of a hydrogen atom are important to understanding the oxidation reactions of MES. The lowest energy molecular structures were identified using the density functional B3LYP/6‐311G(2d,d,p) level of theory. Standard enthalpies of formation (ΔHfo298) for the radicals and their parent molecules were calculated using the density functional B3LYP/6‐31G(d,p), B3LYP/6‐31 + G(2d,p), and the composite CBS‐QB3 ab initio methods. Isodesmic reactions were used to determine ?Hfo values. Internal rotation potential energy diagrams and rotation barriers were investigated using the B3LYP/6‐31G(d,p) level theory. Contributions for So298 and Cp(T) were calculated using the rigid rotor harmonic oscillator approximation based on the structures and vibrational frequencies obtained by the density functional calculations, with contributions from torsion frequencies replaced by internal rotor contributions. The recommended values for enthalpies of formation of the most stable conformers of CH3SCH2CH2, CH2(OOH)SCH2CH3, CH3SCH(OOH)CH3, and CH3SCH2CH2OOH are ?14.0, ?33.0, ?37.2, and ?32.7 kcal/mol, respectively. Group additivity values were developed for estimating properties of structurally similar and larger sulfur‐containing peroxides. Groups for use in group additivity estimation of sulfur peroxide thermochemical properties were developed.  相似文献   

10.
Amine radical cations of the type R3N·+ and [R3NCH2]·+, R=CH3, C3H7, and nitric oxide, NO, have been used to probe the bonding to the surface and the dynamics of the radicals trapped in the confined space of cages or channels in the zeolite. Regular continuous-wave electron spin resonance (ESR) was employed to study the internal motion of the cation radicals formed by γ-irradiation of amines and related ammonium ions, introduced during the synthesis of the zeolites Al-offretite, SAPO-37, SAPO-42 and AlPO4-5. The ESR spectra of [(CH3)3NCH2]·+ radical cation in several studied systems changed reversibly with temperature, indicating dynamical effects. Free rotation about the >N?CH2 bond of the [(CH3)3NCH2]·+ species was found to occur in the temperature range of 110 to 300 K, while the rotation about the >N?CH3 bonds was hindered. The observations confirm the theoretical prediction on the basis of density functional theory calculations, which indicate that the corresponding barriers are of the order of 0.3 and 7 kJ/mol, respectively. The radical cations of the type R3N·+ with R=C2H5, C3H7 were found to undergo a different type of dynamics, involving a two-jump process of the methylene hydrogens next to the nitrogen. A cage or channel size effect on the stability and molecular dynamics was inferred in some cases. Pulsed ESR was employed to study the (NO)2 triplet-state dimers in Na-A type zeolite, with the purpose to resolve the interaction with surface groups, and to elucidate the role of the zeolite on stabilizing the triplet rather than the usual singlet state. Measurements performed at 5 K gave rise to Fourier transform spectra that were assigned to the dimer species interacting with one or more23Na nuclei, with approximative parameters A(23Na)=(4.6, 4.6, 8.2) MHz and Q(23Na)=(?0.3, ?0.3, 0.6) MHz for the hyperfine and nuclear quadrupole coupling tensors, respectively. The values are of similar magnitude as those determined for the NO?Na+ complex. The stability of the triplet-state structure was attributed to unusual geometric structure imposed by the zeolite matrix, with the N?O bonds along a line as in [O?N?Na+?N?O], which according to UHF ab initio calculations has a triplet ground  相似文献   

11.
Recently, we reported the discovery of the first examples of transition metal selenocarbonyl complexes, (n5-C5H5) Mn (CO)2 (CSe) and (n6-C6H5CO2CH2(CO)2 (CSe).1 These complexes are particularly interesting because, unlike CO and CS, the diatomic CSe molecule has so far eluded isolation even at very low temperatures,2 and so these complexes represent the stabilization of a chemically unstable species through coordination to a metal (cf. metal carbene complexes). We have also synthesized several thiocarbonyl complexes of the same type,3,4 as well as the analogous rhenium selenocarbonyl complexes.4 While many detailed studies of the 13C nmr spectra of transition metal carbonyls have appeared in the literature over the past few years, there have been no such studies for the closely related thiocarbonyls and selenocarbonyls, although a few isolated data have been reported recently for metal thiocarbonyls.5,6 In this communication, we report the 13C nmr spectra of the isoelectronic series of complexes, (n5-C5H4R) M(COI)2 (CX) (M = Mn, R = H, CH3; M = Re, R = H; X = O,S, Se).  相似文献   

12.
The temporal variation of chemiluminescence emission from OH?(A2 Σ +) and CH?(A2 Δ) in reacting Ar-diluted H2/O2/CH4, C2H2/O2 and C2H2/N2O mixtures was studied in a shock tube for a wide temperature range at atmospheric pressures and various equivalence ratios. Time-resolved emission measurements were used to evaluate the relative importance of different reaction pathways. The main formation channel for OH? in hydrocarbon combustion was studied with CH4 as benchmark fuel. Three reaction pathways leading to CH? were studied with C2H2 as fuel. Based on well-validated ground-state chemistry models from literature, sub-mechanisms for OH? and CH? were developed. For the main OH?-forming reaction CH+O2=OH?+CO, a rate coefficient of k 2=(8.0±2.6)×1010 cm3?mol?1?s?1 was determined. For CH? formation, best agreement was achieved when incorporating reactions C2+OH=CH?+CO (k 5=2.0×1014 cm3?mol?1?s?1) and C2H+O=CH?+CO (k 6=3.6×1012exp(?10.9 kJ?mol?1/RT) cm3?mol?1?s?1) and neglecting the C2H+O2=CH?+CO2 reaction.  相似文献   

13.
The possibilities of applying three different pulsed ESR techniques have been considered: 1. Phase relaxation measurements by electron spin echo (ESE) affords the estimation of the correlation time of the motion in the region up to 10?5 s. The results of theoretical and experimental analysis of the effect of methyl group rotation in nitroxide radicals have been proposed. 2. The method of pulsed saturation involving detection of ESE signal allows the range of the measured times to be extended up to the values of about 10?2 s. The rotation of CH2 group in the CH2CO2 ? radical and that of the CH3 group in the CH3CHCO2 ? radical have been investigated. 3. The method of pulsed saturation combined with pulsed scanning of H0 allows the analysis of the rotationally induced redistribution of the pulsed saturation throughout the ESR spectra of the radicals. This version of pulsed ESR has been used to study the mobility of nitroxide spin labels.  相似文献   

14.
本文利用266 nm波长的激光及程序升温脱附的方法研究了甲醇在ZnO(0001)表面的光催化反应. TPD结果显示部分的CH3OH以分子的形式吸附在ZnO(0001)表面,而另外一部分在表面发生了解离. 实验过程中探测到H2,CH3·,H2O,CO,CH2O,CO2和CH3OH这些热反应产物. 紫外激光照射实验结果表明光照可以促进CH3OH/CH3O·解离形成CH2O,在程序升温或光照的过程中它又可以转变为HCOO-. CH2OHZn与OHad反应在Zn位点上形成H2O分子. 升温或光照都能促进CH3O·转变为CH3·. 该研究对CH3OH在ZnO(0001)表面的光催化反应机理提供了一个新的见解.  相似文献   

15.
Abstract

The electron spin resonance of gamma and ultraviolet irradiated tetrabutylammonium halides [CH3(CH2)3]4NX (X=I,Cl,Br), tetrabutylammonium hydrogen sulfate [CH3(CH2)3]4NHSO4, tetrabutylammonium periodate [CH3(CH2)3]4NIO4, and ultraviolet photolyzed tetramethylammonium iodide, (CH3)4NI, and tetramethylphosphonium iodide, (CH3)4PI have been investigated between 140 and 350 K. The gamma and ultraviolet irradiation damage centers in tetrabutylammonium compounds were attributed to CH3—CH—CH—CH2, radicals, and ultraviolet photolysis damage centers in tetramethylammonium and phosphonium iodides were attributed to ?H3 radicals. The g values of both radicals are found to be isotropic and g = 2.0030 and 2.0037 respectively to the methylallyl and the methyl radicals. The hyperfine coupling constants of the free electron to the protons in the radicals are reported and discussed.  相似文献   

16.
The addition reaction of CH2OO + H2O CH2(OH)OOH without and with X (X = H2CO3, CH3COOH and HCOOH) and H2O was studied at CCSD(T)/6-311+ G(3df,2dp)//B3LYP/6-311+G(2d,2p) level of theory. Our results show that X can catalyse CH2OO + H2O → CH2(OH)OOH reaction both by increasing the number of rings, and by adding the size of the ring in which ring enlargement by COOH moiety of X inserting into CH2OO···H2O is favourable one. Water-assisted CH2OO + H2O → CH2(OH)OOH can occur by H2O moiety of (H2O)2 or the whole (H2O)2 forming cyclic structure with CH2OO, where the latter form is more favourable. Because the concentration of H2CO3 is unknown, the influence of CH3COOH, HCOOH and H2O were calculated within 0–30 km altitude of the Earth's atmosphere. The results calculated within 0–5 km altitude show that H2O and HCOOH have obvious effect on enhancing the rate with the enhancement factors are, respectively, 62.47%–77.26% and 0.04%–1.76%. Within 5–30 km altitude, HCOOH has obvious effect on enhancing the title rate with the enhancement factor of 2.69%–98.28%. However, compared with the reaction of CH2OO + HCOOH, the rate of CH2OO···H2O + HCOOH is much slower.  相似文献   

17.
We reconsider the principle of the 13C bicarbonate (NaH13CO3) method (13C-BM) for the determination of the CO2 production to obtain an estimate of energy expenditure (EE). Its mathematical concept based on a three-compartmental model is related to the [15N]glycine end product method. The CO2 production calculated by the 13C-BM, RaCO2(13C) is compared to the result from the indirect calorimetry, RCO2(IC). In an interspecies comparison (dog, goat, horse, cattle, children, adult human; body mass ranging from 15 to 350?kg, resting and fasting conditions) we found an excellent correlation between the results of 13C-BM and IC with RCO2(IC)?=?0.703?×?RaCO2(13C), (R2?=?0.99). The slope of this correlation corresponds to the fractional 13C recovery (RF(13C)) of 13C in breath CO2 after administration of NaH13CO3. Significant increase in RF(13C) was found in physically active dogs (0.95?±?0.14; n?=?5) vs. resting dogs (0.71?±?0.10, n?=?17; p?=?.015). The 13C recovery in young bulls was greater in blood CO2 (0.81?±?0.05) vs. breath CO2 (0.73?±?0.05, n?=?12, p?<?.001) and in ponies with oral (0.76?±?0.03, n?=?8) vs. intravenous administration of NaH13CO3 (0.69?±?0.07; n?=?8; p?=?.026). We suggest considering the 13C-BM as a ‘stand-alone’ method to provide information on the total CO2 production as an index of EE.  相似文献   

18.
Thermal effects on ultra-high-molecular-weight polyethylene (UHMWPE) residual radicals during the vitamin E diffusion process were studied in detail. Electron paramagnetic resonance (EPR) technique showed a significant reduction in concentrations of radiation-induced primary (alkyl (–CH2?CH–CH2–), allyl (–CH2?CH=CH–CH–CH2–) and polyenyl (–?CH–[CH=CH–] m –) with m > 3) radicals for both control and vitamin E-doped samples. The concentrations of radiation-induced primary radicals (RIPRs) were found to decrease proportionally with the heat/diffusion time. While the EPR spectra of the control samples showed only polyethylene (PE) radicals, the spectra of vitamin E-doped samples were found to exhibit vitamin E radicals in addition to PE radicals. Of particular interest, the heat involved during vitamin E diffusion plays a significant role in reducing the radiation-induced primary radicals of UHMWPE. For 120 min of heat/diffusion time, the available quantity of primary radicals in control samples were found to be ~7.5 % of initial radicals. The leftover amounts of these primary radicals for vitamin E-doped samples were approximately ~10.0 %. In addition to this, EPR power saturation techniques were also used to assess the effects of initial heat/diffusion treatment on the oxygen-induced residual radicals (OIRRs): R1 (–?CH–[CH=CH–] m –) with m > 3 and R2 (?OCH–[CH=CH–] m –) with m = 2 or 3. It was found that the concentration of OIRRs also decreases proportionally with initial heat/diffusion time. The remaining amount of OIRRs relative to leftover RIPRs after heat/diffusion was found to be approximately 4.0 % in controls and was still found to be 10.0 % in vitamin E-doped UHMWPE. This may indicate that vitamin E slows down the oxidation processes, which may contribute to the strong oxidation resistance of vitamin E-doped UHMWPE.  相似文献   

19.
This paper reported the analysis of dilution effects on the opposed-jet H2/CO syngas diffusion flames. A computational model, OPPDIF coupled with narrowband radiation calculation, was used to study one-dimensional counterflow syngas diffusion flames with fuel side dilution from CO2, H2O and N2. To distinguish the contributing effects from inert, thermal/diffusion, chemical, and radiation effects, five artificial and chemically inert species XH2, XCO, XCO2, XH2O and XN2 with the same physical properties as their counterparts were assumed. By comparing the realistic and hypothetical flames, the individual dilution effects on the syngas flames were revealed. Results show, for equal-molar syngas (H2/CO = 1) at strain rate of 10 s?1, the maximum flame temperature decreases the most by CO2 dilution, followed by H2O and N2. The inert effect, which reduces the chemical reaction rates by behaving as the inert part of mixtures, drops flame temperature the most. The thermal/diffusion effect of N2 and the chemical effect of H2O actually contribute the increase of flame temperature. However, the chemical effect of CO2 and the radiation effect always decreases flame temperature. For flame extinction by adding diluents, CO2 dilution favours flame extinction from all contributing effects, while thermal/diffusion effects of H2O and N2 extend the flammability. Therefore, extinction dilution percentage is the least for CO2. The dilution effects on chemical kinetics are also examined. Due to the inert effect, the reaction rate of R84 (OH+H2 = H+H2O) is decreasing greatly with increasing dilution percentage while R99 (CO+OH→CO2+H) is less affected. When the diluents participate chemically, reaction R99 is promoted and R84 is inhibited with H2O addition, but the trend reverses with CO2 dilution. Besides, the main chain-branching reaction of R38 (H+O2→O+OH) is enhanced by the chemical effect of H2O dilution, but suppressed by CO2 dilution. Relatively, the influences of thermal/diffusion and radiation effects on the reaction kinetics are then small.  相似文献   

20.
Recent theoretical studies have shown that termolecular chemistry can be facilitated through reactions of flame radicals (H, O, and OH) or O2 with highly-energized collision complexes (either radical or stable species) formed in exothermic reactions. In this work, radical-radical recombination reaction induced termolecular chemistry and its impact on combustion modeling was studied. Two recombination reactions, H + CH3 + M → CH4 + M and H + OH + M → H2O + M, were analyzed using ab-initio master equation analyses guided by quasiclassical trajectory results. The dynamics results and the master equation calculations indicate that CH4? and H2O? (formed in the two radical-radical reactions outlined above) react rapidly with flame radicals and O2 at rates that are competitive with collisional cooling. The addition of these processes into conventional combustion modeling requires two modifications: the inclusion of the new nonthermal termolecular reaction rates and the simultaneous reduction of the competing recombination reaction rates. The former is described with newly derived Arrhenius expressions based on quasiclassical trajectories, and the latter is achieved by perturbing the recombination reaction rate during the simulation. Kinetic modeling was used to gauge the impact of including this nonthermal chemistry for H2/CH4-air laminar flames speeds. Inclusion of this nonthermal chemistry has a noticeable impact on simulated flame speeds. The procedure developed here can be utilized to properly quantify the effects of such nonthermal reactions in macroscopic kinetic models.  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