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.
The electrochemical reduction of uracil in dimethyl sulfoxide was investigated, using d.c.and a.c. polarography, cyclic voltammetry, and controlled potential electrolysis. Uracil is reduced in a one-electron step (E1/2=?2.3 V); the apparent number of electrons transferred (n) decreases from one at infinite dilution to one-half at concentrations above 1mM. The concentration dependent n-value is due to proton transfer by the parent compound to the radical anion formed on reduction. Such a proton transfer, which has been observed for 2-hydroxypyrimidine, deactivates part of the uracil, which would otherwise be available for reduction, by formation of the more difficultly reducible conjugate base. The uracil anion forms insoluble mercury salts, producing two oxidation waves (E1/2 of ?0.1 and ?0.3 V); the latter wave is due to formation of a passivating film on the electrode. Digital simulations indicate that the protonation rate exceeds 105M?1 s?1 and that, at low uracil concentration, some of the free radical formed on protonation is further reduced. At concentrations exceeding 1 mM, all of the free radical dimerizes. The effect of added acids and base on the electrochemical behavior is described.  相似文献   

2.
The photoinitiated reactions of 2,4-pyridinedicarbonitrile ( 1 ) and benzophenone in neutral, acidic, and basic 3:1 2-propanol-water and the kinetics of disappearance of 1 , have been studied. Pyridinyl radical anion forms as an intermediate by an electron transfer. In acidic solution substitution of the cyano group in the 2 position and in the 2 and 4 positions with diphenylmethanol occurs. In neutral medium both substitution at the 2-position and reduction, in which the cyano group at the 4 position is replaced by hydrogen, are observed. In basic solution in which protonation of the radical anion is not likely, only reduction occurs. The rates of formation and relative yields of these products depends on the pH of the solution. A mechanism is discussed.  相似文献   

3.
Species produced by electron transfer to variously substituted bicyclo[3.1.0]hex-3-en-2-ones provide a better insight into the origin of 1,6-addition products sometimes observed by reaction of lithium diorganocuprates with β-cyclopropyl α-enones. Cyclic voltammetry of eight such bicyclohexenones show that the half-lives of the corresponding anion radicals are very short (t12?10?4s) except when the initial molecule is phenyl substituted at C-4. In such cases the anion radicals are very stable (t12?6s) owing to the greater charge delocalization and we observe a second wave corresponding to the formation of the Dianion. The reduction of the same substrates by the solvated electron in liquid ammonia exhibits the same difference of behaviour. The molecules giving strongly reactive anion radicals only lead to the straight-forward product expected by conjugate reduction; while the 4-phenyl substituted substrates yield a mixture of products from normal conjugate reduction and rearrangement. This correlation strongly suggest that these last products arise from the rearrangement of the dianion.  相似文献   

4.
《Electroanalysis》2003,15(1):19-25
The electrochemical behavior of 2‐(5‐amino‐ 1,3,4‐oxadiazolyl)‐5‐nitrofuran (NF359) and its comparison with well‐known drugs such as nifurtimox (NFX) and nitrofurazone (NFZ) in protic, mixed and aprotic media by cyclic voltammetry, tast and differential pulse polarography was studied. All the compounds were electrochemically reducible in all media being the reduction of the nitrofuran group the main voltammetric signal. The one‐electron reduction couple due to the nitro radical anion formation was visualized in mixed (for NF359 and NFZ) and aprotic media (for all compounds). By applying a cyclic voltammetric methodology we have calculated the decay constants (k2) of the corresponding nitro radical anions in mixed and aprotic media. In mixed medium data fit well with a disproportionation reaction of the nitro radical anion but in aprotic medium fit better with a dimerization reaction. Also, considering cyclic voltammetric measurements in aprotic media we have estimated the reduction potential of the RNO2/RNO2.? couple in aqueous medium, pH 7 (E17 values) finding very good correlation with E17 values obtained by pulse radiolysis. Furthermore we have calculated the equilibrium constants from the electron transfer from nitro radical anion to oxygen (kO2) finding that nitro radical anion from NF359 is thermodynamically favored to react with oxygen in respect to both NFZ and NFX.  相似文献   

5.
Study of oxidation, protonation, and alkylation of the products obtained by one- and two-electron reduction of p-tolunitrile with sodium in liquid ammonia showed that these products are, respectively, p-tolunitrile radical anion and 1-cyano-4-methyl-2,5-cyclohexadienyl anion. The latter is formed by protonation of p-tolunitrile dianion with ammonia. The two-electron reduction involves protonation of p-tolunitrile dianion with initial p-tolunitrile, which gives rise to 4-cyanobenzyl anion. The yield of the latter depends on the order of mixing of the reactants. The anionic reduction products react with butyl bromide to afford products corresponding to ipso alkylation with respect to the cyano group, 4-butyltoluene and 3-butyl-3-cyano-6-methyl-1,4-cyclohexadiene, as well as the alkylation product at the benzylic position, 4-pentylbenzonitrile. The formation of 4-butyltoluene indicates the possibility for selective synthesis of p-alkyltoluenes by reductive alkylation of p-tolunitrile.  相似文献   

6.
Irradiation of the α-fluoroketones 1a and 6a in i-PrOH selectively affords the parent ketone 1b and 6b , respectively. It is concluded that in this solvent heterolytic C-F bond cleavage of the anion radical-formed by electron transfer to the excited fluoroketone-is a faster process than the subsequent protonation by the cation radical of the solvent. In cyclohexane 1b and 6b are only formed in minor amounts, the fluorinated RH-reduction product 4 now being the major product from 1a . In non-reducing solvents as t-BuOH or benzene 2-fluorocyclohexanone (1a) exhibits a similar behavior as cyclohexanone (1b) on excitation. The quantum yields for α-cleavage are alike for both compounds, but oxetane formation with 2-methylpropene as olefinic component occurs much more readily with 1a than with 1b .  相似文献   

7.
Despite the fundamental importance of radical‐anion radical‐cation pairs in single‐electron transfer (SET) reactions, such species are still very rare and transient in nature. Since diborenes have highly electron‐rich B? B double bonds, which makes them strong neutral reductants, we envisaged a possible realization of a boron‐centered radical‐anion radical‐cation pair by SET from a diborene to a borole species, which are known to form stable radical anions upon one‐electron reduction. However, since the reduction potentials of all know diborenes (E1/2=?1.05/?1.55 V) were not sufficiently negative to reduce MesBC4Ph4 (E1/2=?1.69 V), a suitable diborene, IiPr?(iPr)B?B(iPr)?IiPr, was tailor‐made to comply with these requirements. With a halfwave potential of E1/2=?1.95 V, this diborene ranks amongst the most powerful neutral organic reductants known and readily reacted with MesBC4Ph4 by SET to afford a stable boron‐centered radical‐anion radical‐cation pair.  相似文献   

8.
The first main‐group element radical based one‐dimensional magnetic chain ( 1K )n was realized by one‐electron reduction of the pyridinyl functionalized borane 1 with elemental potassium in THF in the absence of 18‐crown‐6 (18‐c‐6). The electron spin density of ( 1K )n mainly resides at the boron centers with a considerable contribution from central benzene and pyridine moieties. The spin centers exhibit an antiferromagnetic interaction as demonstrated by magnetic measurements and theoretical calculations. In contrast, the reduction in the presence of 18‐c‐6 afforded the separated radical anion salt 1K(Crown) , in which the potassium cation was trapped by THF and 18‐c‐6 molecules. Further one‐electron reduction of 1K(Crown) and ( 1K )n led to the diamagnetic monomer and polymer, respectively.  相似文献   

9.
It is shown that the first reversible one-electron wave in the reduction of 5-arylfurans in anhydrous dimethylformamide (DMF) corresponds to the formation of an anion radical and that the subsequent waves are associated with cleavage of the C-Hal bond in the case of halo derivatives and with reduction of the anion radical and the arylfuran fragment. The character of the reduction of 5-(p-nitrophenyl)furan derivatives is determined by the ability of the substituent in the 2 position to delocalize the negative charge. In conformity with this, the first two reversible waves of carbonyl-substituted derivatives are one-electron waves and correspond to the formation of a stable dianion, the greatest contribution to the resonance hybrid of which is made by a p-quinoid structure. The second wave of 5-(p-nitrophenyl)furan and its 2-CH2OH derivative is irreversible and corresponds to the transfer of three electrons. Lithium ions have a substantial effect on the height and E1/2 value of the second reduction wave, and this effect is manifested more markedly, the less the substituent in the 2 position is capable of delocalizing the negative charge. The transmission factor of the furan ring is 0.48.Translated from Khimiya Geterotsiklicheskikh Soedinenii, No. 2, pp. 165–172, February, 1977.  相似文献   

10.
The electrochemical behaviour of temozolomide on a glassy carbon electrode has been investigated. The reduction of temozolomide is an irreversible process, pH dependent, and the mechanism involves the addition of one electron and one proton to C5 to form an anion radical, causing the irreversible breakdown of the tetrazinone ring. The oxidation mechanism of temozolomide is an irreversible, adsorption‐controlled process, pH dependent up to value close to the pKa and occurs in two consecutive charge transfer reactions, with the formation of the hydroxylated product. The electroanalytical determination of TMZ led to a detection limit of 1.1 µM.  相似文献   

11.
The heteroaromatic polynitrile compound tetracyanopyridine (TCNPy) is introduced as a new electron acceptor for the formation of deeply colored charge‐transfer complexes. In MeCN, TCNPy is characterized by a quasireversible one‐electron‐reduction process at ?0.51 V (versus SCE). The tetracyanopyridine radical anion undergoes a secondary chemical reaction, which is assigned to a protonation step. TCNPy has been demonstrated to generate 1:1 complexes with various electron donors, including tetrathiafulvalene (TTF) and dihydroxybenzene derivatives, such as p‐hydroquinone and catechol. Visible‐ or NIR‐light‐induced excitation of the intense charge‐transfer bands of these compounds leads to a direct optical electron‐transfer process for the formation of the corresponding radical‐ion pairs. The presence of available electron donors that contain protic groups in close proximity to the TCNPy acceptor site opens up a new strategy for the photocontrolled generation of pyridinium radicals in a stepwise proton‐coupled electron‐transfer (PCET) sequence.  相似文献   

12.
Abstract— Laser flash photolysis has been used to investigate the mechanism of formation and decay of the radical species generated by light-induced electron transfer from chlorophyll a (Chi) triplet to various quinones in egg phosphatidyl choline bilayer vesicles. Chlorophyll triplet quenching by quinone is controlled by diffusion occurring within the bilayer membrane (kq~ 106M?1 s?1. as compared to ~ 109 M?1 s?1 in ethanol) and reflects bilayer viscosity. Radical formation via separation of the intermediate ion pair is also inhibited by increased bilayer viscosity. Cooperativity is observed in the radical formation process due to an enhancement of radical separation by electron transfer from semiquinone anion radical to a neighboring quinone molecule. Two modes of radical decay are observed, a rapid (t1/2= 150μ) recombination between Chi and quinone radicals occurring within the bilayer and a much slower (t1/2= 1–100 ms) recombination occurring across the bilayer-water interface. The latter is also cooperative, which accounts for a t1/2 which is dependent upon quinone concentration. The slow decay is only observed with quinones which are not tightly anchored into the bilayer, and is probably the result of electron transfer from semiquinone anion radical formed within the bilayer to a quinone molecule residing at the bilayer-water interface. Direct evidence for such a process has been obtained from experiments in which both ubiquinone and benzoquinone are present simultaneously. With benzo-quinone, approx. 60% of the radical decay occurs via the slow mode. Triplet to radical conversion efficiencies in the bilayer systems are comparable to those obtained in fluid solution (~ 60%). However, radical recombination, at least for the slow decay mechanism, is considerably retarded.  相似文献   

13.
Single-electron oxidation in acetonitrile and reduction in DMF of sulfides of 3-RS-1,4-dimethyl-7- ethylazulenes (R = Me, Et, Ph, p-MeC6H4, p-MeOC6H4, N-1-phenyltetrazolyl) leads to stable radical cations and radical anions, respectively. The found reduction potentials of sulfides of the azulene series are close to those of natural and synthetic bioantioxidants. In the radical cations the unpaired electron is essentially delocalized over the cyclopentadienyl fragment of the molecule, and the sulfide group in large measure defines the distribution of spin density. In the radical anion spin density is delocalized in the tropylium cycle, and the influence of the sulfide group is insignificant. Electrochemical oxidation of unsubstituted 1,4-dimethyl-7-ethylazulene results in the formation of a dimeric radical cation.  相似文献   

14.
Five pathways leading to the deamination of cytosine (to uracil) after formation of its deprotonated radical cation are investigated in the gas phase, at the UB3LYP/6‐311G(d,p) level of theory, and in bulk aqueous solvent. The most favorable pathway involves hydrogen‐atom transfer from a water molecule to the N3 nitrogen of the deprotonated radical cation, followed by addition of the resulting hydroxyl radical to the C4 carbon of the cytosine derivative. Following protonation of the amino group (N4), the C4? N4 bond is broken with elimination of the NH3?+ radical and formation of a protonated uracil. The rate‐determining step of this mechanism is hydrogen‐atom transfer from a water molecule to the cytosine derivative. The associated free energy barrier is 70.2 kJ mol?1.  相似文献   

15.
Metalation of 6,13,20,21‐tetraaryl‐22H‐[14]tribenzotriphyrins(2.1.1) (TriP, 1 a – d ) with [Mn(CO)5Br] provided MnI tricarbonyl complexes of [14]tribenzotriphyrins(2.1.1) 2 a – d in 85–93 % yield. The complexes were characterized by mass spectrometry and UV/Vis absorption, IR, and NMR spectroscopy. Single‐crystal X‐ray analyses revealed that 2 b and 2 c adopt bowl‐shaped conformations. The redox properties of [(TriP)MnI(CO)3] ( 2 a – d ) were studied by cyclic voltammetry. Each compound undergoes two reversible one‐electron reductions to form a porphyrin π anion radical and a dianion in CH2Cl2. Two oxidation waves were observed, the first of which corresponds to a metal‐centered electron‐transfer process. The redox potentials of 2 a – d are consistent with the optical spectroscopic data and the relatively narrow HOMO–LUMO gaps that were predicted in DFT calculations. The optical spectra can be assigned by using Michl’s perimeter model. TDDFT calculations predict the presence of several metal‐to‐ligand charge‐transfer bands in the L‐band region between 500 and 700 nm.  相似文献   

16.
Chemical single‐electron reduction of 1‐mesityl‐2,3,4,5‐tetraphenylborole ( 3 ) gave a stable radical anion [CoCp*2][ 3 ] as shown in earlier investigations. Herein, we present the reaction of [CoCp*2][ 3 ] with the 2,2,6,6‐tetramethylpiperidine‐N‐oxyl radical (TEMPO), a common radical trap. Instead of radical recombination, the reaction proceeds through a redox pathway involving oxidation of the borole radical anion combined with reduction of TEMPO. This electron‐transfer process is accompanied by a deprotonation reaction of the cobaltocenium counterion by the base TEMPO? to give TEMPO‐H and a neutral cobalt(I) fulvene complex ( 7 ). The latter was not observed directly during the reaction, because it instantaneously reacts as a nucleophile attacking at the boron center of the in situ generated borole 3 to give the borate 6 . However, 7 was synthesized independently by deprotonation of [CoCp*2][PF6]. In addition, the obtained zwitterionic cobaltocenium borate 6 undergoes a photolytic rearrangement to form the borata‐alkene derivative 9 that thermally transforms to the chiral cobaltocenium borate 12 . Our investigations are based on spectroscopic evidence, X‐ray crystallography, elemental analysis, as well as DFT calculations.  相似文献   

17.
Photochemical transformations of phenothiazine (PTA) in solutions of halomethanes CHnX4–n (X = Cl, Br; n = 0, 1, 2) and in n-hexane—CHnX4–n mixtures under the irradiation with = 337 and 365 nm were studied. The rate constants of quenching of PTA fluorescence with halomethanes (k q) are 4·105—1.3·1010 L mol–1 s–1. The process occurs due to electron transfer with the C—X bond cleavage in the radical anion fragment of the primary radical ion pair. This results in the formation of the stable radical cation salt (PTA·+X). The plot of k q vs. free energy of electron transfer corresponds to the Rehm—Weller empirical equation for a one-electron process and is satisfactorily described in terms of the theory of nonradiative electron transitions in the approximation of one quantum vibration.  相似文献   

18.
Beware of samarium diiodide and aryl ketones! If the ketyl radical anion which is formed by electron transfer finds a properly placed aryl group, a highly diastereoselective cyclization may occur. After the transfer of a second electron and protonation bi- and polycyclic products with a common 1,4-cyclohexadiene moiety may be isolated [Eq. (a)]. X=CHCO2R, NCH2Ph; HMPA=(Me2N)3PO.  相似文献   

19.
Kinetics of formation and recombination of radicals formed by quenching of the triplet state of 4-carboxybenzophenone (CB) with para-substituted phenol derivatives RC6H4OH (R = OMe, H, Cl, Br, I) in aqueous solutions was studied by nanosecond laser photolysis. At pH ≥ 5.4, quenching proceeds with high rate constants ((1–3)⋅109 L mol−1 s−1) through electron transfer to form the radical anion CB⋅− and radical cation RC6H4OH⋅+. The latter is transformed into the phenoxyl radical within ≤10 ns. At pH ≤ 8, the CB⋅− radical anion is protonated in a phosphate buffer with the rate constant increasing from 4⋅106 to 15⋅106 s−1 with a decrease in the pH from 8 to 5.4. The yield of radicals decreases from 100 to 13% as the atomic weight of halogen in the RC6H4OH molecule increases due to an increase in the probability of recombination of the primary triplet radical pair in the solvent cage and partial intersystem crossing in an encounter complex (3CB, RC6H4OH). The effect of heavy atom is also observed in the kinetics of volume recombination of the radicals, the magnitude of effect corresponds to the acceleration of the primary recombination of the triplet radical pair. Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 6, pp. 1397–1402, June, 2005.  相似文献   

20.
Thanks to the superior redox potential of platinum(II) complex compared with that of Ru(bpy)32+ in the excited state, an efficient and selective visible‐light‐induced CDC reaction has been developed by using a catalytic amount (0.25 %) of 1 . With the aid of FeSO4 (2 equiv), the corresponding amide could not be detected under visible‐light irradiation (λ=450 nm), but the desired cross‐coupling product was exclusively obtained under ambient air conditions. A spectroscopic study and product analysis revealed that the CDC reaction is initiated by photoinduced electron‐transfer from N‐phenyltetrahydroisoquinoline to the complex. An EPR (electron paramagnetic resonance) experiment provides direct evidence on the generation of superoxide radical anion (O2? . ) rather than singlet oxygen (1O2) under irradiation of the reaction system, in contrast to that reported in the literature. Combined, the photoinduced electron‐transfer and subsequent formation of superoxide radical anion (O2? . ) results in a clean and facile transformation.  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