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 93 毫秒
1.
The evolution of the hyperfine tensors in the nitroxide series with increasing alkyl substitution on the NO group has been studied theoretically for radicals from H2NO to C5H10NO. A projection technique has been applied to the UHF wave-functions in order to correct spin densities for quartet contamination. The magnitude of the isotropic and anisotropic couplings reflects the substitution effect already observed on the spin distribution maps, that is a spin transfer from oxygen to nitrogen when alkyl groups are substituted to hydrogens in H2NO. The alternation of the signs of the couplings along the chain as well as the cos2 γ law for the coupling constants of atoms (C or H) in β position are verified in the series. The orientation of the anisotropic tensors with respect to the chemical bonds depends on the position of the atoms in the molecule (radical site, α, β positions); it is not affected by further substitution, except for strongly asymmetric configurations.  相似文献   

2.
Molar extinction coefficients of aqueous solutions of some organic compounds, viz. formamide (CH3NO),N-methylformamide (C2H5NO),NN-dimethylformamide (C3H7NO),NN-dimethylacetamide (C4H9NO), 1,4-dioxane (C4H8O24), succinimide (C4H5NO2) and solutions of acetamide (C2H5NO) and benzoic acid (C7H6O2) in 1,4-dioxane (C4H8O2) have been determined by narrow beam γ-ray transmission method at 81, 356, 511, 662, 1173 and 1332 keV. The experimental values of mass attenuation coefficients of these compounds have been used to calculate effective atomic numbers and electron densities. The additivity rule earlier used for aqueous solution has been extended to non-aqueous (1,4-dioxane) solutions.  相似文献   

3.
The Jahn-Teller (JT) and the E.P.R. parameters of the trigonal, JT-active hexakispyridine-N-oxide complex Cu(C5H5NO)6 2- are calculated from the results of extended Hückel molecular orbital calculations on a series of distorted geometries of this complex. These geometries are obtained from a recent, low temperature, neutron diffraction study and compared with a monoclinic, static, modification of the corresponding nitrate complex. For two different distortion models, the JT minimum is obtained for a tetragonally elongated octahedron. The JT parameters are in very reasonable agreement with the experimental ones. The EPR parameters are compared with newly determined experimental values obtained by a re-evaluation, using matrix diagonalization instead of a perturbation treatment, of earlier measurements on Cu(C5H5NO)6(BF4)2 at 4·2 K and on Cu(C5H5NO)5DMF(ClO4 2 at room temperature. On the basis of these computations, the E.P.R. parameters, and especially the very small isotropic copper hyperfine coupling, can be understood.  相似文献   

4.
ENDOR spectra of a series of carbon- and oxygen-centered radical adducts of 2-substituted DMPO-type nitrones are reported. They include the novel cyclic nitrones, 2-phenyl-5,5-dimethyl-pyrroline-N-oxide (2-Ph-5,5-M2PO), 2,5,5-trimethyl-l-pyrroline-N-oxide (2,5,5-M3PO), and 2-phenyl-3,3,5,5,-tetramethyl-l-pyrroline-N-oxide (2-Ph-3,3,5,5-M4PO). Electron paramagnetic resonance (EPR) was used to ascertain the nitrogen hyperfine splittings (hfs’s) while1H ENDOR was employed to determine the long-range (γ) hydrogen hfs’s. The magnitude of the nitrogen hfs combined with the numbers and sizes of the long-range γ-H hfs’s of spin adducts of these new spin traps are shown to help disclose the identities of various added radicals (or radical addends). It should be noted that the three new spin traps presented here are keto-nitrones not aldo-nitrones. Thus, there is alkyl (e.g. CH3) or aryl (e.g. C6H5) substitution at the 2-position of the pyrroline-N-oxide ring. This feature is part of our search for modified spin traps that yield spin adducts with greater stability.  相似文献   

5.
The luminescence quenching of excited Tris(2,2-bipyridine)ruthenium(II) ions by trans-[RuCl2{P(OR)3}4] complexes with different alkyl chain ligands (R=C2H5, C2H5Cl, nC4H9, iC3H7 o-tolyl and tC4H9) was investigated. None of the acceptor Ru(II) phosphite complexes were luminescent, and the rate constants of the bimolecular system were determined within the range of 1.15 and 0.28×108 M−1 s−1 for R=C2H5 and tC4H9, respectively. The results indicate a direct effect of the alkyl chains in the rate constants, showing a decrease of kq as a function of increased of the alkyl chains (R) in the ruthenium(II) tetraphosphite complexes. The greater the R group content in the phosphite ligand, the more difficult the electron transfer is.  相似文献   

6.
《Composite Interfaces》2013,20(7):617-627
In order to control the surface wettability of hyperbranched hydrophilic poly(amidoamine) (PAMAM)-grafted nano-sized silica, hydrophilic alkyl chain (C n H2n+1) with different chain lengths (n = 4, 8, 15) were postgrafted onto PAMAM-grafted silica by the reaction of terminal amino groups of PAMAM grafted on the silica surface with alkyl acid chlorides (C n H2n+1-COCl). The postgrafting of C n H2n+1-COCl increased with increasing PAMAM grafting and alkyl chain length of C n H2n+1-COCl. However, the terminal amino groups of PAMAM-grafted silica used for the postgrafting of C n H2n+1-COCl decreased with increasing chain length. This may be due to the steric hindrance between terminal amino groups of PAMAM-grafted silica and C n H2n+1-COCl: the steric hindrance is considered to increase with increasing chain length of C n H2n+1-COCl. The surface wettability was estimated by contact angle measurement for water and methanol wettability. As a result, it was found that contact angle and methanol wettability increased with increasing alkyl chain length of postgrafted C n H2n+1-COCl. The hyperbranched PAMAM-grafted silica readily dispersed in water and methanol because of the hydrophilic nature of grafted PAMAM, but it lost dispersibility in water and methanol due to postgrafting of hydrophobic chains.  相似文献   

7.
Yuhai Hu  Keith Griffiths   《Surface science》2008,602(17):2949-2954
Fourier transform infra red reflection–absorption spectroscopy (FTIR-RAS), thermal desorption spectroscopy (TDS), and auger electron spectroscopy (AES), were employed to explore the mechanism of NO reduction in the presence of C2H4 on the surface of stepped Pt(3 3 2). Both NO–Pt and C2H4–Pt interactions are enhanced when NO and C2H4 are co-adsorbed on Pt(3 3 2). As a result, C2H4 is dissociated at surface temperatures as low as 150 K, and the N–O stretch band is weakened. The presence of post-exposed C2H4 leads NO desorption from steps to decrease significantly, but the same effect on NO desorption from terraces becomes appreciable only at higher post-exposures of C2H4, e.g., 0.6 L and 1.2 L, and proceeds to a much slighter extent. Auger spectra indicate that as a result of the reaction with O from NO dissociation, the amount of surface C species is greatly reduced when NO is post-exposed to a C2H4 adlayer. It is concluded that reduction of NO in the presence of C2H4 proceeds very effectively on the surface of the Pt(3 3 2), through a mechanism of NO dissociation and subsequent O removal. Following this mechanism, the significant dissociation of adsorbed NO molecules on steps at surface temperatures below 400 K, and subsequent rapid reaction between the resultant O and C-related species, accounts for the considerable amount of N2 desorption at temperatures below 400 K.  相似文献   

8.
The nature and strength of metal–ligand bonds in organotransition–metal complexes is crucial to the understanding of organometallic reactions and catalysis. The Fe‐N homolytic bond dissociation energies [ΔHhomo(Fe‐N)′s] of two series of para‐substituted Fp anilines p‐G‐C6H4NHFp [1] and p‐G‐C6H4N(COMe)Fp [2] were studied using the Hartree–Fock (HF) and the density functional theory methods with large basis sets. In this study, Fp is (η5‐C5H5)Fe(CO)2 and G are NO2, CN, COMe, CO2Me, CF3, Br, Cl, F, H, Me, MeO and NMe2. The results show that BP86 and TPSSTPSS can provide the best price/performance ratio and accurate predictions of ΔHhomo(Fe‐N)′s. B3LYP can also satisfactorily predict the α and remote substituent effects on ΔHhomo(Fe‐N)′s [ΔΔHhomo(Fe‐N)′s]. The good correlations [r = 0.96 (g, 1), 0.99(g, 2)] of ΔΔHhomo(Fe‐N)′s in series 1 and 2 with the substituent σp+ constants imply that the para‐substituent effects on ΔHhomo(Fe‐N)′s originate mainly from polar effects, but those on radical stability originate from both spin delocalization and polar effects. ΔΔHhomo(Fe‐N)′s(1,2) conform to the captodative principle. Insight from this work may help the design of more effective catalytic processes. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

9.
The bay-capping mechanism on PAH armchair edges and the kinetics of acetylene addition to 6–6–5 and 5–6–5 bays have been explored by ab initio/RRKM-ME calculations. The bays on the edges were modeled by C21H11 and C20H9 radicals produced by H abstractions from 7H-benzo[c]cyclopenta[e]pyrene and dicyclopenta[cf]pyrene. The C20H9 + C2H2 reaction is shown to have a low entrance barrier and to rapidly form the capped product, indaceno[2,1,8,7-cdefg]pyrene, along with ethynyl substituted dicyclopenta[cf]pyrene at temperatures above 1400 K. The reactivity of C21H11 is shown to be governed by the location of the unpaired electron; the π radical R1 formed by H abstraction from the CH2 group in 7H-benzo[c]cyclopenta[e]pyrene reacts with C2H2 very slowly owing to a high entrance barrier, with the bay-capping rate constant approaching 10−16 cm3 molecule−1 s−1 only at temperatures above 2000 K. This result reaffirms that the growth of π aryl radicals via acetylene addition is inefficient and reflects the generally low reactivity of such radicals where the spin density is highly delocalized over the entire polyaromatic system. Alternatively, the σ C21H11 radical R2 produced by H abstraction from the five-membered ring at the bay rapidly reacts with C2H2 forming the bay-capped product, with the rate constant on the order of 10−12 cm3 molecule−1 s−1 at T ≥ 1500 K. Rate constants for the capping reactions at the 6–6–5 and 5–6–5 bays are compared with those at the 6–0–6, 6–6–6, and 6–5–6 bays. The site-specific bay-capping rate constants have been utilized in kMC simulations of the PAH growth and the results showed measurable differences when the 6–6–5 and 5–6–5 bay-capping reactions are taken into account, including an increase of the growth rate and the formation of closed-shell PAH and a rise of the number of embedded five-membered rings accompanied with a slight decrease of their overall amount.  相似文献   

10.
Methods of determining rate data for elementary reactions for combustion applications, using experimental and theoretical methods, are briefly reviewed. The approaches are illustrated by reference to recent research in three areas: (i) reactions of OH with C2H4 and C2H2, where theory, tuned by reference to experiment, has provided a substantial contribution to the determination of rate data for these complex reactions, over a wide range of temperatures; (ii) reactions between alkyl radicals and O2, where theory and experiment have been closely allied in discerning the details of mechanisms for small alkyl radicals; much remains to be done with larger radicals; (iii) reactions of methylene and the interactions between chemical reaction and the conversion of the singlet state into the triplet, where theory has played little part thus far. Comments are also made on the process of evaluating rate data for elementary reactions for incorporation in chemical mechanisms for use in combustion models.  相似文献   

11.
A two-sublattice X-Y antiferromagnet is treated using the second-order Green function theory. The higher-order Green functions are decoupled by the random phase approximation (RPA), and the transition temperature TN, where the paramagnetic staggered susceptibility diverges, and the critical values of the spin correlation functions are calculated for systems having simple cubic and body centred cubic lattice structures and for all spin values. Comparing the results with the experimental Néel temperatures of Co(C5H5NO)6(ClO4)2 and Co(C5H5NO)6(BF4)2 we get J/kB=0.258 K and 0.216 K for these compounds, respectively.  相似文献   

12.
Hydrophobic forms of the N,N‐dialkyl‐4‐nitroaniline (DNAP) (p‐O2NC6H4NR2) ( 1a–f ) and alkyl‐4‐nitrophenyl ether (p‐O2NC6H4OR) ( 2a–c ) solvatochromic π* indicators have been characterized and compared with respect to: (a) solvatochromic bandshape, (b) sensitivity expressed as ?s , ( / d π * ), and (c) trends in ? s with increasing length of alkyl chain(s) on the probe molecule. ? Octyl 4‐nitrophenyl ether (p‐O2NC6H4OC8H17) ( 2b ) and ? decyl 4‐nitrophenyl ether (p‐O2N C6H4 OC10H21) ( 2c ) were synthesized and their solvatochromic UV/Vis absorption bands were found to maintain a Gausso‐Lorentzian bandshape for the indicators in non‐polar and alkyl substituted aromatic solvents, for example, hexane(s) and mesitylene. Corresponding absorption bands for 1a–f display increasing deviation from a Gausso‐Lorentzian shape in the same solvents as the alkyl chains on the indicator are increased in length all the way to C10 and C12, for example, N,N‐didecyl‐4‐nitroaniline (p‐O2NC6H4N (C10H21)2) and N,N‐didodecyl‐4‐nitroaniline (p‐O2NC6H4N (C12H25)2) ( 1d–f ). A plot of ? s versus Cn follows a 1st order decay for the DNAP indicators but is linear for the alkyl 4‐nitrophenyl ethers. A discussion of how the long alkyl chains on the two types of indicators affect the orientation and overlap of n and π * orbitals, and resulting solvatochromic bands is presented. For DNAP, overextending the alkyl chains to obtain greater hydrophobic character may cause the alkane component to dominate solute‐solvation processes at the expense of the probe's fundamental solvatochromic character. Copyright © 2007 John Wiley & Sons, Ltd.  相似文献   

13.
The radicals formed by radiation in 2-thiobarbituric acid single crystals have been studied by E.S.R. spectroscopy. The main radical is formed by abstraction of one of the protons bonded to C5. The electronic structure of this radical has been compared to that of the related radical formed in ordinary barbituric acid. Both the experimental data as well as the MO calculations show that when oxygen is substituted by sulphur the isotropic proton splitting is reduced and the g-value anisotropy increased. This is explained by a transfer of unpaired spin density from C5 to the sulphur atom.  相似文献   

14.
Hydrogenated boron-carbon clusters, i.e. organo-borons, have received considerable attention both theoretically and experimentally. Herein, using a topology searching strategy, we systematically explore the structures and stabilities of small organo-borons with CBHz (z ≤ 8) and CxB3-xHz (x = 1, 2,z≤ 14) stoichiometry, with particular interests in the intrinsic stabilities of the organo-boron radicals. At the CCSD(T)/aug-cc-pVQZ//B3LYP/aug-cc-pVTZ and CCSD(T)/aug-cc-pVQZ//MP2/aug-cc-pVTZ levels, the stabilities of these global minimum organo-boron species were evaluated by considering dissociation pathways and the binding energy per atom. Aside from the five already studied radicals (CBH2, CBH4, C2BH2, C2BH4 and CB2H3), we predict six novel radicals, i.e. CBH6, C2BH6, C2BH8, CB2H, CB2H5 and CB2H7, which could be detected under suitable circumstances. However, observation of the highly hydrogenated CB2H9 radical is much less likely due to its minute stability towards H-extrusion. The computationally determined stable/meta-stable maximum hydrogenation numbers for CB, C2B and CB2 (6, 8 and 8, respectively) are in excellent agreement with a simple electron-counting model for CxBy chains. The newly predicted organo-boron radicals await future laboratory verification.  相似文献   

15.
A laser-induced fluorescence (LIF) system was developed to diagnose the reaction and transport of radicals in the ArF-laser-assisted CVD environment. The C2(a 3 II u) radicals were produced by the multi-photon dissociation of C2H2. The transport of the radical was directly measured by LIF. The disappearance rate of the radical in C2H2 was also determined. By using the values determined, the in-flux of the C2 radicals onto a substrate was investigated, based on a diffusion model.  相似文献   

16.
Transport of exhausted thermonuclear fuel in the ITER divertor and pumping duct was modeled on a specially designed dc glow discharge setup using mass spectrometry, optical and electron microscopy, and electron probe microanalysis. Transport and deposition of hydrocarbon radicals transferred in an H2/C x H yx mixture through a hollow stainless steel anode at a total mixture pressure of 8–212 Pa and a methane content to 15 mol % were considered. It was shown that deposition of radicals and ions (CH3, C2H3, C2H5) with kinetic energies of 0.03–3 eV on the anode inner surface at 600 K was suppressed to a large extent. In the temperature range of 600–800 K, deposition of ions and radicals with kinetic energy of ~3 eV was partially restored with the formation of soft a-C:H films, while thermalized radicals were not condensed.  相似文献   

17.
The resonantly-stabilized cyclopentadienyl radical (C5H5) is a key species in the combustion and molecular growth kinetics of mono and poly-aromatic hydrocarbons (M/PAHs). At intermediate-to-low temperatures, the C5H5 reaction with the hydroperoxyl radical (HO2) strongly impacts the competition between oxidation to smaller products and growth to PAHs, precursors of soot. However, literature estimates for the HO2 + C5H5 reaction rate are inaccurate and inconsistent with recent theoretical calculations, thus generating discrepancies in global combustion kinetic models. In this work, we perform state-of-the-art theoretical calculations for the HO2 + C5H5 reaction including variable reaction coordinate transition state theory for barrierless channels, accurate thermochemistry, and multi-well master equation (ME) simulations. Contrary to previous studies, we predict that OH + 1,3-C5H5O is the main reaction channel. The new rate constants are introduced in two literature kinetic models exploiting our recently developed ME based lumping methodology and used to perform kinetic simulations of experimental data of MAHs oxidation. It is found that the resonantly-stabilized 1,3-C5H5O radical is the main C5H5O isomer, accumulating in relevant concentration in the system, and that the adopted lumping procedure is fully consistent with results obtained with detailed kinetics. The reactivity of C5H5O with OH and O2 radicals is included in the kinetic mechanisms based on analogy rules. As a result, C5H5O mostly reacts with O2 producing smaller C3/C4 species and large amounts of C5H4O, suggesting that further investigations of the reactivity of both C5H5O and C5H4O with oxygenated radicals is necessary. Overall, this work presents new reliable rate constants for the HO2 + C5H5 reaction and provides indications for future investigations of relevant reactions in the sub-mechanisms of cyclopentadiene and MAH oxidation.  相似文献   

18.
The finite-temperature (T) properties of the muonium substituted ethyl radical CH2MuCH2 have been theoretically studied by Feynman path integral quantum Monte Carlo (PIMC) simulations. To derive the ensemble averaged expectation values we have combined the PIMC formalism with an efficient tight-binding (TB) Hamiltonian and a density functional operator of the B3LYP type in the EPRIII basis. The TB operator has been used to calculate the potential energy surface (PES) of the ethyl radical in the doublet ground state, the harmonic and anharmonic vibrational wave numbers as well as several probability density functions of the nuclei. The harmonic linear response approximation, which makes use of the Feynman centroid density, has been adopted to evaluate the anharmonic wave numbers. The large anharmonicities in the nuclear potential lead to bond lengths in thermal equilibrium which exceed the vibrationless parameters at the PES minimum. This enhancement is particularly strong for the C–Mu bond. It is responsible for the suppression of the intramolecular rotation for temperatures below room temperature. In C2 H5 the rotation is allowed down to 10?K. The dissimilar rotational dynamics for H2MuCH2 and C2 H5 has been studied with the help of TB-based probability density functions. The nuclear configurations of CH2MuCH2 and C2 H5, which are populated in thermal equilibrium, have been used to evaluate the isotropic and anisotropic hyperfine splitting (hfs) constants under explicit consideration of the nuclear vibrations and the internal rotation. The hfs constants have been determined with the help of the B3LYP-EPRIII Hamiltonian. The hindered low-temperature rotation in the Mu isomer is responsible for roto-vibrational corrections to the isotropic hfs constants which are smaller than the corrections in C2 H5. The shortcomings of single-configuration approaches for the evaluation of isotropic hfs constants have been demonstrated for both radicals. The ensemble corrections to the isotropic hfs parameters are correlated with fluctuations in the atomic spin densities. Differences in the absolute values of the isotropic hfs parameters in CH2MuCH2 and C2 H5 can be traced back to differences in the nuclear degrees of freedom. The ensemble shift for each isotropic hfs parameter can be explained by characteristic nuclear motions. For this discussion we make use of the distribution functions of the isotropic hfs constants. Roto-vibrational corrections to the anisotropic hfs constants are rather small. PIMC simulations have been performed between 25 and 1000?K, i.e. in a T interval that is large enough to consider nuclear effects beyond zero-point motions. The TB and B3LYP-EPRIII based physical quantities of CH2MuCH2 and C2 H5 have been compared with experimental findings whenever possible.  相似文献   

19.
A novel hidden reaction of the phenoxyl radical (C6H5O?) with a specific daughter is found to significantly alter its hitherto accepted coupling reactions' scheme. Transient characterizations and mechanistic evaluations in highly acidic to strongly alkaline aqueous medium reveal this concurrent reaction competing favorably in nanosecond–microsecond time‐scale with the five distinct C6H5O? + C6H5O? reactions, which produce various phenolic end‐products as reported earlier (M. Ye and R. H. Schuler, J. Phys. Chem. 1989, 93, 1898). Presently, only the symmetric 4,4′‐dioxo transient precursor, O?C6H5? H5C6?O that leads to the stable 4,4′‐biphenol product, gets partially oxidized by a fraction of remaining C6H5O?. The resulting secondary transient ?C12H9O2 radical is generated at diffusion‐controlled rate, k > 5.0 × 109 M?1 s?1, and follows an independent chemistry. Consequently, when the previously reported five coupled end product distribution ratios were appropriately updated, the respective fractional values revealed a closer match for the symmetric 2,2′‐ and 4,4′‐biphenols with their suggested coupling reaction branching probabilities based on the atomic spin‐density distributions in the C6H5O? radical (P. Neta, R. W. Fessenden, J. Phys. Chem., 1974, 78, 523). Results also suggest that in the remaining fraction, differential solvation in aqueous medium of various orientation‐related encounter complexes (C6H5O…C6H5O) formed during coupling favors rearrangement only toward 2,4′‐biphenolic product, at the cost of 2‐ and 4‐phenoxyphenolic species. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

20.
13C-E.S.R. coupling constants of triphenylmethyl radical are readily obtained from the model compound tris-(3,5-di-t-butylphenyl)-methyl. Twisting of the phenyl substituents in solution is proven for the first time by the observation of a considerable contribution from hyperconjugation to acortho C. This twisting causes the unusually large a meta H by π-σ interaction as proposed by Pople for twisted benzyl radicals. Two different sets of spin densities are obtained for C1 and C2 by employing ac2 C, coupling constants of the ring protons and , depending on whether the σ-π parameters of Karplus and Fraenkel or those recently proposed by Strom, Underwood and Jurkowitz are used. The comparison of ac1 C calculated from these spin densities with the experimental carbon-13 splitting at C1 indicates some deviation from trigonal planarity at C1 in solution when the σ-π-parameters of Karplus and Fraenkel are used. Use of the Strom, Underwood and Jurkowitz parameters, on the other hand, lead to pure sp2 hybridization at C1. The latter result is in contrast to the deviations from trigonal planarity observed previously for the central carbon atoms of triphenylmethyl radical in the gas phase and of tris-(p-nitrophenyl)-methyl in the solid state.  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