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 15 毫秒
1.
The substructure associated with the doublet ESR spectrum of the acetaldehyde radical cation in Freon matrices below 100 K is shown to arise from a matrix interaction and not, as previously proposed, from coupling to the hydrogens of the methyl group. Since the reversible loss of this substructure at higher temperature is accompanied by almost no change in the large isotropic 1H coupling(136 G) to the aldehydic hydrogen, the matrix perturbation appears to have a negligible effect on the spin distribution in the radical cation and is according interpreted as a superhyperfine interaction  相似文献   

2.
It has been established that, upon X-ray irradiation of various methyloxiranes in freonic matrices at 77 K, both open and cyclic (with the elongated C-C bond) forms of radical cations are stabilized. It has been shown that observed reversible photoinduced transformations of 2,3-dimethyloxirane and methyloxirane radical cations are related to the conversion between the open and cyclic forms of the radical cations with high quantum yields (0.02?C0.39, depending on the oxirane and the matrix). For the trimethyloxirane radical cation the action of light on the trans-isomer of the open form results in its photoinduced transformation into a C-centered radical with low quantum efficiency (??4 × 10?3). Tetramethyloxirane radical cations, stabilized in their open form, are resistant to the action of light. Probable causes of the observed effects are discussed. Upon the X-ray irradiation of 2,2-dimethyloxirane in freonic matrices at 77 K, a cyclic form of the radical cation is stabilized (presumably, as part of a complex with matrix molecules) which transforms into a distonic C-centered radical cation under the action of light with the quantum yield of ??10?3.  相似文献   

3.
A 4 K matrix ESR study shows that the molecular radical cations of isopropyl formate and acetate, produced radiolytically in halocarbon matrices at 4.2 K, undergo spontaneous rearrangement due to a selective intramolecular hydrogen shift from the tertiary CH bond in the isopropyl group to the carbonyl oxygen atom giving RC+(OH)OC(CH3)2, where R = H or CH3. The radical cation of tert-butyl acetate undergoes further fragmentation at the ester CO bond following a similar rearrangement to give an isobutene radical cation in CFCl3.  相似文献   

4.
In this work, we regiospecifically generate and compare the gas-phase properties of two isomeric forms of tryptophan radical cations—a distonic indolyl N-radical (H3N+ - TrpN?) and a canonical aromatic π (Trp?+) radical cation. The distonic radical cation was generated by nitrosylating the indole nitrogen of tryptophan in solution followed by collision-induced dissociation (CID) of the resulting protonated N-nitroso tryptophan. The π-radical cation was produced via CID of the ternary [CuII(terpy)(Trp)] ?2+ complex. CID spectra of the two isomeric species were found to be very different, suggesting no interconversion between the isomers. In gas-phase ion-molecule reactions, the distonic radical cation was unreactive towards n-propylsulfide, whereas the π radical cation reacted by hydrogen atom abstraction. DFT calculations revealed that the distonic indolyl radical cation is about 82 kJ/mol higher in energy than the π radical cation of tryptophan. The low reactivity of the distonic nitrogen radical cation was explained by spin delocalization of the radical over the aromatic ring and the remote, localized charge (at the amino nitrogen). The lack of interconversion between the isomers under both trapping and CID conditions was explained by the high rearrangement barrier of ca.137 kJ/mol. Finally, the two isomers were characterized by infrared multiple-photon dissociation (IRMPD) spectroscopy in the ~1000–1800 cm–1 region. It was found that some of the main experimental IR features overlap between the two species, making their distinction by IRMPD spectroscopy in this region problematic. In addition, DFT theoretical calculations showed that the IR spectra are strongly conformation-dependent.   相似文献   

5.
The unimolecular dissociation reaction of the methyl vinyl ether radical cation to acetyl cation plus methyl radical was studied by means of 13C labelling and tandem mass spectrometric techniques. The results are in good agreement, in all major aspects, with previous data for 13C-labeled propene oxide. The data are discussed in terms of non-ergodic behaviour. Isomerization of the methoxyethylidene radical cation intermediate to the acetone radical cation is ruled out on the basis of the labelling experiments and comparison with the hydroxymethene radical cation system.  相似文献   

6.
The radical cations of linear alkanes (n-pentane, n-heptane) trapped in various matrices (Freon-11, Freon-113, Freon-113a, mixture of Freon-11 and Freon-114B2, and sulfur hexafluoride) were found to undergo the following types of photochemical reactions: (1) charge transfer to the matrix followed by neutralization, (2) isomerization and unimolecular decomposition, and (3) deprotonation. The absorption spectra of the radical cations were characterized, and the quantum yields of reactions occurring in different matrices at 77 K were determined. It was shown that the reaction pathway and efficiency of the photochemical processes observed for a given radical cation in different matrices with similar physical and chemical characteristics could considerably differ.  相似文献   

7.
The thermal and photochemical reactions of the methylal radical cation /I/ in freon matrices were studied using selective deuteration for elucidating the structure of the resulting species. /I/ has been shown to decay by unimolecular reaction upon heating to 140 K as well as upon photolysis in CFCl3 matrix and the product of decay has been assumed to be the complex of formaldehyde radical cation with CFCl3. Such decay reaction has been demonstrated for 1,3-dioxolan radical cation as well.  相似文献   

8.
The intrinsic chemical properties of the gaseous adenine radical cation were examined by using dual cell Fourier transform ion cyclotron resonance mass spectrometry. The adiabatic recombination energy of the radical cation (ionization energy of neutral adenine) was found by bracketing experiments to be 8.55 ± 0.1 eV (at 298 K; earlier literature values range from 8.3 to 8.9 eV). Based on this value, the heat of formation (ΔHf298) of the adenine radical cation is estimated to be 246 ± 3 kcal/mol. The acidity (ΔHacid298) of the adenine radical cation was bracketed to be 221 ± 2 kcal/mol. These thermochemical values suggest that the adenine radical cation reacts with neutral guanine by electron abstraction or proton transfer, with neutral cytosine by proton transfer, and via neither pathway with neutral thymine, molecular water or a sugar moiety of DNA (modeled by tetrahydrofuran). Experimental examination of the gas-phase reactivity of the adenine radical cation revealed a slow deuterium atom abstraction from perdeuterated tetrahydrofuran. Hence, in the absence of a nearby guanine or cytosine, the adenine radical cation may be able to abstract a hydrogen atom from a sugar moiety of DNA.  相似文献   

9.
Initial processes of radiation-induced cationic polymerization of styrene and α-methylstyrene have been studied by means of microsecond pulse radiolysis. For styrene, absorption bands caused by the monomer cation radical St+? appear at 630 and 350 nm in a mixture of isopentane and n-butyl chloride at about ?165°C. In parallel with the decay of St+?, three absorption bands appear in the near-infrared (IR) region, and at 600 and 450 nm. The IR and 600 nm bands are assigned to the associated dimer cation radical St2+?, and the 450 nm band to the bonded dimer cation radical St-St+?. The kinetic behavior of these species shows that reaction of St+? with styrene monomer forms both St2+? and St-St+?. With the decay of St-St+?, another absorption band appears at 340 nm, and the lifetime of this band is relatively long. The 340 nm band may be due to carbonium ions of the growing polystyrene. For α-methylstyrene, the monomer cation radical (at 690 and 350 nm), the associated dimer cation radical (in the near-IR region and at 620 nm) and the bonded dimer cation radical (at 480 nm) behave in a manner similar to that of the corresponding styrene species. The absorption band caused by carbonium ions of growing poly(α-methylstyrene) appears at 340 nm.  相似文献   

10.
The structure of the methyl tert-butyl ether (MTBE) radical cation and mechanism of its thermal and photochemical reactions in irradiated freons (CFCl3, CF2ClCFCl2, and CF3CCl3) were studied. Radical products of MTBE radiolysis in the liquid phase were investigated by the spin trapping technique. The quantum-chemical calculations of the structure of MTBE radical cations and products of their transformations were carried out by density functional theory (DFT) and ab initioMP2 methods. The primary MTBE radical cations are stabilized in dilute solutions in CFCl3and CF3CCl3. The ion–molecule reaction (proton transfer from the radical cation) was found to occur in concentrated solutions in CFCl3immediately during irradiation. The action of light ( = 436 to 546 nm) at 77 K on the MTBE radical cation in CFCl3and CF3CCl3matrices results in intramolecular migration of the methyl group to yield the distonic radical cation (CH3)2 .CO+(CH3)2. The primary MTBE radical cations undergo an irreversible transformation with methane elimination resulting in formation of the 2-methoxypropene radical cation .CH2=+(3)3in CFCl3and CF3CCl3matrices in the temperature range 110–130 K. In the case of CF2ClCFCl2matrix, such a reaction occurs during irradiation at 77 K. Using the spin trapping technique, it was shown that the liquid-phase radiolysis of the neat ether resulted in the formation of fragmentation products (.CH3,CH3., and t-BuO. radicals) from the primary radical cations, as well as the products of their rearrangements and ion–molecule reactions.  相似文献   

11.
p-Methoxystyrene (MOS), butyl vinyl ether (BVE), and N-vinylcarbazole (VCZ) were polymerized in high yield by azoinitiators such as 2, 2'-azodiisobutyronitrile (AIBN) in the presence of electron acceptors such as Ph2I+PF6. An electron paramagnetic resonance (ESR) study of the model radicals of the propagating radical showed the transformation of the radical to the corresponding cation in the presence of the electron acceptors. In the case of BVE, the polymer formation was caused by cationic species produced by the transformation of the initiating radical. The polymerizations of MOS and VCZ were ascribed to the transformation of the growing radical to the corresponding cation during the propagation step which was classified as the radical/cation transformation polymerization. Block copolymers of MOS/cyclohexene oxide (1, 2-epoxycyclohexane) (CHO) and VCZ/CHO were effectively prepared by the radical/cation transformation polymerization of the appropriate monomers in the presence of AIBN, electron acceptor and CHO. The formation of block copolymers was characterized by turbidimetry, thin-layer chromatography, and solubility tests.  相似文献   

12.
The photochemistry of 2,6‐dimethyl‐4‐chlorophenol ( 6 ) has been studied in methanol and trifluoroethanol (TFE) through product studies and transient absorption spectroscopy. Chloride loss from triplet 6 gave triplet hydroxyphenyl cation 14 , which equilibrated with triplet oxocyclohexadienylydene 15 within a few tens of nanoseconds; the cation can, however, be selectively trapped by allyltrimethylsilane (kad = 108–109 m ?1 s?1) to give a phenonium ion and the allylated phenol. In neat alcohols, 14 and 15 are reduced through different mechanisms, namely by hydrogen transfer through radical cation 17 and via phenoxyl radical 16 , respectively. The mechanistic rationalization has been substantiated by the parallel study of an Osilylated derivative. The work shows that the chemistry of the highly (but selectively) reactive phenyl cation 14 can not only be discriminated from that of the likewise highly reactive carbene 15 , but also exploited for synthetically useful reactions, as in this case with alkenes. Photolysis of electron‐donating substituted halobenzenes may be the method of choice for the mild generation of some classes of phenyl cations.  相似文献   

13.
It was found that the principal photochemical reaction of 1,3,5-trioxane radical cations in freonic matrices at 77 K is their cycle-opening dissociation yielding the distonic radical cation in which the unpaired electron is preferentially localized on the oxygen atom. The dissociation of the trioxane radical cations at 77 K is characterized by high quantum yields, which vary from 0.24 to 0.36 in different matrices. The distonic radical cations produced during photolysis are unstable at 77 K and undergo further transformations, which occur at different rates in freonic matrices. The structure of the intermediates produced and a possible mechanism of the processes are discussed with the use of quantum-chemical calculation data.  相似文献   

14.
The addition of 1 and 2 molar equivalents of bromine to a series of 10-alkylphenothiazines, 1a-d (methyl, ethyl, n-propyl, and isopropyl, respectively), yields the corresponding 3-bromo- and 3,7-dibromo-10-alkylphenothiazines ( 11a-d and 12a-d , respectively). Evidence which supports the typical clectrophilic aromatic substitution mechanism is presented. Radical cations ( 12a-d.+ ) arc produced when 12a-d are treated with 1 or 2 molar equivalents of bromine. Upon boiling in acetic acid these radical cations are converted predominantly to 1,3,7,9-lelrabromophenothiazine ( 5 ) and the parent 3,7-dibromo-10-alkylphenothiazine ( 12a-d ) with the evolution of hydrogen bromide. The 10-methyl radical ( 12a ) gives, in addition, 1,3,7-tribromo-10-methylphenothiazine ( 15 ). A mechanism if proposed for these reactions in which initial dealkylution of 12b-d.+ to 3.7-dibromophenothiazine radical cation ( 13 ) occurs followed by reduction of 13.+ by bromide ion to parent 3,7-dibromophenothiazine ( 13 ). Subsequent bromination of 13 by molecular bromine produced in the previous redox reaction yields 1,3,7-tribromo-( 14 ) and 1,3,7,9-tetra-bromo-( 5 ) phenothiazines. The small size of the methyl group allows 12a to be brominated at the 1-position prior to dealkylation. In addition to undergoing bromination at the 3- and 7-position, 10-isopropylphenothiazine ( 1d ) is oxidized to the radical cation 12e.+ when treated with bromine. 10-Benzylphenothiazine ( 1e ), however, undergoes oxidation to radical cation 1e.+ exclusively. This radical cation debenzylates readily at room temperature and is converted finally into phenothiazine.  相似文献   

15.
Radical cation formation is proposed for the rapid cyclization of 1, 2-bis[5-phenyl-2-methylthien-3-yl]cyclopentene and oligothiophene functionalized dimethyldihydropyrenes (DMDHP). Density functional theory calculations have been performed to rationalize the effect of a radical cation on the activation barrier of different classes of electrocyclic photochromes (DHP, dithienylethene, dihydroazulene and fulgide). For exact comparative analysis, the activation barrier of neutral (singlet) analogues at the same level of theory are also calculated. In addition, the concerted nature and aromaticity of transition states were investigated with the help of synchronicity (Sy.) and nuclear independent chemical shift values NICS(0) calculations, respectively, for both the radical cation and neutral systems. In case of the radical cation, thermal return of CPD to DHP, the activation barrier is very low (ΔH = 3.13 kcal mol?1, ΔG = 4.01 kcal mol?1) as compared to the neutral analogue (ΔH = 20.6 kcal mol?1, ΔG = 20.98 kcal mol?1), which is consistent with experimental observations. Similarly for dithenylethenes, radical cation formation has a large impact on the activation barrier (ΔH = 19.44 kcal mol?1, ΔG = 22.29 kcal mol?1). However, radical cation formation has almost negligible impact on the activation barrier of VHF-DHA and fulgide isomerization. The significant difference has been observed for synchronicity and NICS(0) values of all types of photochromes under radical cation conditions as compared to the neutral system.  相似文献   

16.
The adsorption of DPH in M6.6ZSM-5 (M=Na+, K+, Rb+, Cs+), RbFER and RbMOR channel zeolites takes place without chemical or structural modification. After photoexcitation of these systems, a radical cation–electron pair is observed and has a sufficiently long lifetime to be studied by diffuse reflectance UV-visible spectroscopy. The study of the recombination of this radical cation-electron pair was carried out at different temperatures and allowed the determination of the activation energy as a function of the nature of the charge-balancing cation but also of the confinement effect. It appears that the activation energy decreases progressively from Na+ to Cs+ but also when the confinement decreases. To go further, the free enthalpies have been calculated from the Marcus theory demonstrating experimentally that these systems are located in the inverted Marcus region.  相似文献   

17.
para‐Phenylene‐bridged spirobi(triarylamine) dimer 2 , in which π conjugation through four redox‐active triarylamine subunits is partially segregated by the unique perpendicular conformation, was prepared and characterized by structural, electrochemical, and spectroscopic methods. Quantum chemical calculations (DFT and CASSCF) predicted that the frontier molecular orbitals of 2 are virtually fourfold degenerate, so that the oxidized states of 2 can give intriguing electronic and magnetic properties. In fact, the continuous‐wave ESR spectroscopy of radical cation 2 .+ showed that the unpaired electron was trapped in the inner two redox‐active dianisylamine subunits, and moreover was fully delocalized over them. Magnetic susceptibility measurements and pulsed ESR spectroscopy of the isolated salts of 2 , which can be prepared by treatment with SbCl5, revealed that the generated tetracation 2 4+ decomposed mainly into a mixture of 1) a decomposed tetra(radical cation) consisting of a tri(radical cation) moiety and a trianisylamine radical cation moiety (≈75 %) and 2) a diamagnetic quinoid dication in a tetraanisyl‐p‐phenylendiamine moiety and two trianisylamine radical cation moieties (≈25 %). Furthermore, the spin‐quartet state of the tri(radical cation) moiety in the decomposed tetra(radical cation) was found to be in the ground state lying 30 cal mol?1 below the competing spin‐doublet state.  相似文献   

18.
The electrochemical and chemical oxidation (by hinderedo-benzoquinones or NOClO4) of H2S in nonaqueous solutions (MeCN) proceeds with the donation of one electron. The formation of the unstable radical cation of hydrogen sulfide was detected by cyclic voltammetry. The radical cation decomposes to form H+ and the HS. radical. The generation of the hydrogen sulfide radical cation was confirmed by ESR spectroscopy in a frozen Freon matrix. The possibility of using the hydrogen sulfide radical cation in the synthesis of organosulfur compounds under mild conditions was studied. The concept of the work was proposed by Prof. O. Yu. Okhlobystin. The first electrochemical experiments were performed when he was alive. Published inIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 7, pp. 1182–1188, July, 2000.  相似文献   

19.
Kinetic and mechanistic studies on the formation of an oxoiron(IV) porphyrin cation radical bearing a thiolate group as proximal ligand are reported. The SR complex, a functional enzyme mimic of P450, was oxidized in peroxo‐shunt reactions under different experimental conditions with variation of solvent, temperature, and identity and excess of oxidant in the presence of different organic substrates. Through the application of a low‐temperature rapid‐scan stopped‐flow technique, the reactive intermediates in the SR catalytic cycle, such as the initially formed SR acylperoxoiron(III) complex and the SR high‐valent iron(IV) porphyrin cation radical complex [( SR .+)FeIV?O], were successfully identified and kinetically characterized. The oxidation of the SR complex under catalytic conditions provided direct spectroscopic information on the reactivity of [( SR .+)FeIV?O] towards the oxidation of selected organic substrates. Because the catalytically active species is a synthetic oxoiron(IV) porphyrin cation radical bearing a thiolate proximal group, the effect of the strong electron donor ligand on the formation and reactivity/stability of the SR high‐valent iron species is addressed and discussed in the light of the reactivity pattern observed in substrate oxygenation reactions catalyzed by native P450 enzyme systems.  相似文献   

20.
This study demonstrates how positive ion atmospheric pressure photoionization-ion mobility spectrometry-mass spectrometry (APPI-IMS-MS) can be used to produce different ionic forms of an analy te and how these can be separated. When hexane:toluene (9:1) is used as a solvent, 2,6-di-tert-butylpyridine (2,6-DtBPyr) and 2,6-di-tert-4-methylpyridine (2,6-DtB-4-MPyr) efficiently produce radical cations [M]+ and protonated [M + H]+ molecules, whereas, when the sample solvent is hexane, protonated molecules are mainly formed. Interestingly, radical cations drift slower in the drift tube than the protonated molecules. It was observed that an oxygen adduct ion, [M + O2]+, which was clearly seen in the mass spectra for hexane:toluene (9:1) solutions, shares the same mobility with radical cations, [M]+. Therefore, the observed mobility order is most likely explained by oxygen adduct formation, i.e., the radical cation forrning a heavier adduct. For pyridine and 2-tert-butylpyridine, only protonated molecules could be efficiently formed in the conditions used. For 1- and 2-naphthol it was observed that in hexane the protonated molecule typically had a higher intensity than the radical cation, whereas in hexane:toluene (9:1) the radical cation [M]+ typically had a higher intensity than the protonated molecule [M + H]+. Interestingly, the latter drifts slower than the radical cation [M]+, which is the opposite of the drift pattern seen for 2,6-DtBPyr and 2,6-DtB-4-MPyr.  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