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 459 毫秒
1.
The chemical processes of tryptophan (Trp), tyrosine (Tyr) and a dipeptide Trp-Tyr, which are induced by UV radiation and one-electron oxidation of SO. 4 , have been investigated in aqueous solution by KrF (248 nm) laser flash photolysis. On the basis of optical studies, the photoionization of Trp and Tyr produces the tryptophan indolyl radical and tyrosine phenoxyl radical, respectively, and these are different from the intermediates resulting from interaction of Trp and Tyr with SO. 4 . In the case of Trp, SO. 4 would attack the indole moiety to produce a C(2)-yl sulphate radical adduct, and Tyr is oxidized to produce mainly the corresponding one-electron oxidized radical, which deprotonates rapidly to form the phenoxyl radical in neutral solution, and a possible sulphate radical adduct. From transient absorption spectra of photoionization of Trp-Tyr, an intramolecular electron transfer, Trp/N.-Tyr Trp-Tyr/O., has been observed, but there was no observation of the process of one-electron oxidation of Trp-Tyr by SO. 4 .  相似文献   

2.
Abstract— Flash photolysis transient spectra (Λ > 250 nm) of aqueous papain show that the initial products are the neutral tryptophan radical Tip (Λmax 510 nm), the tryptophan triplet state 3Trp (Λm., 460nm), the disulfide bridge electron adduct –?S-— (Λmax 420nm) and the hydrated electron eaq-. The –?S-– yield was not altered by nitrous oxide or air, indicating that the formation of this product does not involve electrons in the external medium. The original papain preparation was activated by irradiating under nitrogen. The action spectrum supports previous work attributing the low initial activity to blocking of cysteinyl site 25 with a mixed disulfide. Flash lamp irradiation in nitrogen led to activation at low starting activities and inactivation at higher starting activities, while only inactivation at the same quantum yield was observed with air saturation. The results are consistent with photoionization of an essential tryptophyl residue as the key inactivating step.  相似文献   

3.
There has been considerable interest in the photochemistry of tryptophan in connection with ultraviolet inactivation of enzymes. Earlier flash photolysis work has demonstrated that the hydrated electron (e-aq) is an initial product in the irradiation of indole derivatives, accompanied by a longer-lived transient absorption near 500 nm attributed to an aromatic radical species[1–5]. Similar transients were observed in a recent flash photolysis study of lysozyme[6] in which it was proposed that inactivation is a consequence of electron ejection from 1 to 2 essential tryptophan residues in the active center. However, there has been uncertainty concerning the tryptophan radical structure and its relationship to the triplet state and radical spectra reported for tryptophan photolysis in low-temperature rigid media. This note reports a flash photolysis investigation of L-tryptophan (Trp) and 1-Methyl-L-tryptophan (1-MeTrp) undertaken to clarify these points. The flash photolysis apparatus and methods employed are described in Ref. [6].  相似文献   

4.
355 nm光照下利用瞬态吸收光谱技术进行了有氧、无氧条件下二苯醚与亚硝酸体系的反应机理研究, 考察了其中瞬态物种的衰减行为, 并对其光解产物进行了GC-MS分析. 研究表明, HNO2在355 nm紫外光的照射下产生的OH自由基和二苯醚反应生成C12H10O-OH 加合物, N2条件下C12H10O-OH衰减的速率常数为(1.86±0.14)×105 s-1, 在有氧条件下, C12H10O-OH可转化为C12H10O-OHO2, 衰减的速率常数为(6.6±0.4)×106 s-1. N2条件下最终产物为苯酚、2-羟基二苯醚、4-羟基二苯醚、4-硝基二苯醚.  相似文献   

5.
The relationship between peptide structure and electron transfer dissociation (ETD) is important for structural analysis by mass spectrometry. In the present study, the formation, structure and reactivity of the reaction intermediate in the ETD process were examined using a quadrupole ion trap mass spectrometer equipped with an electrospray ionization source. ETD product ions of zwitterionic tryptophan (Trp) and Trp‐containing dipeptides (Trp‐Gly and Gly‐Trp) were detected without reionization using non‐covalent analyte complexes with Ca2+ and 18‐crown‐6 (18C6). In the collision‐induced dissociation, NH3 loss was the main dissociation pathway, and loss related to the dissociation of the carboxyl group was not observed. This indicated that Trp and its dipeptides on Ca2+(18C6) adopted a zwitterionic structure with an NH3+ group and bonded to Ca2+(18C6) through the COO? group. Hydrogen atom loss observed in the ETD spectra indicated that intermolecular electron transfer from a molecular anion to the NH3+ group formed a hypervalent ammonium radical, R‐NH3, as a reaction intermediate, which was unstable and dissociated rapidly through N–H bond cleavage. In addition, N–Cα bond cleavage forming the z1 ion was observed in the ETD spectra of Trp‐GlyCa2+(18C6) and Gly‐TrpCa2+(18C6). This dissociation was induced by transfer of a hydrogen atom in the cluster formed via an N–H bond cleavage of the hypervalent ammonium radical and was in competition with the hydrogen atom loss. The results showed that a hypervalent radical intermediate, forming a delocalized hydrogen atom, contributes to the backbone cleavages of peptides in ETD. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

6.
The Rose Bengal‐sensitized photooxidations of the dipeptides l ‐tryptophyl‐l ‐phenylalanine (Trp‐Phe), l ‐tryptophyl‐l ‐tyrosine (Trp‐Tyr) and l ‐tryptophyl‐l ‐tryptophan (Trp‐Trp) have been studied in pH 7 water solution using static photolysis and time‐resolved methods. Kinetic results indicate that the tryptophan (Trp) moiety interacts with singlet molecular oxygen (O2(1Δg)) both through chemical reaction and through physical quenching, and that the photooxidations can be compared with those of equimolecular mixtures of the corresponding free amino acids, with minimum, if any, influence of the peptide bond on the chemical reaction. This is not a common behavior in other di‐ and polypeptides of photooxidizable amino acids. The ratio between chemical (kr) and overall (kt) rate constants for the interaction O2(1Δg)‐dipeptide indicates that Trp‐Phe and Trp‐Trp are good candidates to suffer photodynamic action, with krlkt values of 0.72 and 0.60, respectively (0.65 for free Trp). In the case of Trp‐Tyr, a lower krlkt value (0.18) has been found, likely as a result of the high component of physical deactivation of O2(1Δg) by the tyrosine moiety. The analysis of the photooxidation products shows that the main target for O2(1Δg) attack is the Trp group and suggests a much lower accumulation of kynurenine‐type products, as compared with free Trp. This is possibly because of the occurrence of another accepted alternative pathway of oxidation that gives rise to 3a‐oxidized hydrogenated pyrrolo[2,3‐b]indoles.  相似文献   

7.
Aerobic and anaerobic photolysis of methyl(pyridine)cobaloxime, benzyl(pyridine)cobaloxime and analogous compounds in CHCl3 results only in an electron transfer reaction from an equatorial ligand producing photo-reduction of CoIII to CoII, the complex retaining its axial ligands.If after the anaerobic photolysis of benzyl(pyridine)cobaloxime the oxygen is introduced without any further photolysis we obtain an ESR spectrum of nitroxide, arising from the attack of a benzyl radical on the dimethylglyoxime equatorial ligand.For the other complexes, homolytic cleavage of the CoC bond occurs and in the presence of oxygen gives rise to the superoxide cobalt complex adduct Py(CoIIIO2?.During photolysis of methyl(pyridine)cobaloxime in isopropanol homolytic cleavage of the CoC bond occurs in preference to electron transfer reaction from the equatorial ligands.The anaerobic photolysis of benzyl(pyridine)cobaloxime in isopropanol or in water at 113–133 K results in an electron transfer reaction. However, at 170 K we observe the formation of the CoII complex arising from CoC bond cleavage.A mechanism for photo-induced insertion of oxygen in the CoC bond is proposed.  相似文献   

8.
The photolysis of anhydrous zinc azide prepared by (1) dehydration of Zn(N3)2·2H2O and (2) precipitation by acetone from aqueous solution, under the action of high pressure mercury arc is reported here. The pressure of nitrogen developed during photolysis under steady-state conditions at constant intensity is a linear function of t1/2 in both the samples. The rate of photolysis is a linear function of intensity at constant temperature. A detailed analysis of the dark rate suggests that the dependence on t1/2 should be due to the diffusion of nitrogen from the reaction site to the surface of the solid azide. An appropriate mathematical analysis is presented. The linear dependence of the rate of photolysis on the intensity of irradiation is interpreted in terms of the reaction of a trapped exciton with an adjacent azide ion. Absence of any detectable photoconduction in the wavelength range of the irradiating radiation suggests that excitons are the most probable reactive species. The sample of zinc azide obtained by the first method decomposes faster than the other under identical conditions, and has a slightly lower energy of activation of 1.4 kcal/mole compared to 1.75 kcal/mole for the second. These differences are discussed in terms of the defect concentrations of the two azide samples.  相似文献   

9.
Reactions of carbonate radical (CO3 ), generated by photolysis or by radiolysis of a carbonate solution with nickel(II)-iminodiacetate (Ni(II)IDA) were studied at pH 10.5 and ionic strength (I)==0.2 mol·dm–3. The stable product arising from the ligand degradation in the complex is mainly glyxalic acid. Time-resolved spectroscopy and transient kinetics were studied using flash photolysis. From the kinetic data it was suggested that the carbonate radical initially reacts with Ni(III)IDA with a rate constant (2.4±0.4)·106 dm3·mol–1·s–1 to form a Ni(II)IDA species which, however, undergoes a first-order transformation (k=2.7·102·s–1) to give a radical intermediate of the type Ni(II)RNHCHCO 2 ) which rapidly forms an adduct containing a Ni–C bond. This adduct decays very slowly to give rise to glyoxalic acid. From a consideration of equilibrium between Ni(II)IDA and Ni(III)IDA, the one electron reduction potential for the Ni(III)IDA/Ni(II)IDA couple was determined to be 1.467 V.  相似文献   

10.
Photoinduced hydroxylation of neat deaerated benzene to phenol occurred under visible‐light irradiation of 2,3‐dichloro‐5,6‐dicyano‐p‐benzoquinone (DDQ), which acts as a super photooxidant in the presence of water. Photocatalytic solvent‐free hydroxylation of benzene derivatives with electron‐withdrawing substituents such as benzonitrile, nitrobenzene, and trifluoromethylbenzene used as neat solvents has been achieved for the first time by using DDQ as a super photooxidant to yield the corresponding phenol derivatives and 2,3‐dichloro‐5,6‐dicyanohydroquinone (DDQH2) in the presence of water under deaerated conditions. In the presence of dioxygen and tert‐butyl nitrite, the photocatalytic hydroxylation of neat benzene occurred with DDQ as a photocatalyst to produce phenol. The photocatalytic reactions are initiated by oxidation of benzene derivatives with the singlet and triplet excited states of DDQ to form the corresponding radical cations, which associate with benzene derivatives to produce the dimer radical cations, which were detected by the femto‐ and nanosecond laser flash photolysis measurements to clarify the photocatalytic reaction mechanisms. Radical cations of benzene derivatives react with water to yield the OH‐adduct radicals. On the other hand, DDQ . ? produced by the photoinduced electron transfer from benzene derivatives reacts with the OH‐adduct radicals to yield the corresponding phenol derivatives and DDQH2. DDQ is recovered by the reaction of DDQH2 with tert‐butyl nitrite when DDQ acts as a photocatalyst for the hydroxylation of benzene derivatives by dioxygen.  相似文献   

11.
Abstract— A pyrimidine adduct, 6-4‘-[pyrimidine-2’-one] thymine (PO-T)?, observed in DNA hydrolysates of 254-nm ultraviolet (u.v.) irradiated conidia of Streptomyces coelicolor, increases linearly with u.v. dose up to 2 × 105 ergs/mm2. Yields of thymine dimer (T○) and uracil-thymine dimer (U○) level off at much lower doses. Initial relative rates of formation of these u.v. photoproducts are: 1:1.3:4.8 for PO-T, T○ and U○, respectively. Similar results were obtained with a Streptomyces griseus mutant, PHR-1. An equation is derived to estimate the ratio of the amount of PO-T to the total amount of thymine-derived photoproducts at low (biological) u.v. doses. The observed PO-T fractions compare well with the calculated values. Rapid photolysis of the precursor of PO-T was observed by post-u. v. treatment at 313 nm of conidia of S. coelicolor and of S. griseus PHR-1. The photolysis was much slower at 365 nm and did not occur at all at 405 nm. Pyrimidine dimers were not appreciably affected by post-u. v. treatment at the above wavelengths in these Streptomyces strains. Both of these strains are phenotypically photoreactivation-deficient, and the present results indicate that they do not possess active photoreactivating enzyme. In earlier papers[3,4,5], the pyrimidine adduct found in acid hydrolysates of DNA was loosely referred to as “uracil-thymine adduct (U-T adduct)”. Such terminology is not strictly correct. The pyrimidine adduct in acid hydrolysates is PO-T (sometimes called P2B), which could theoretically result from removal of ammonia from a C-T adduct or removal of water from a U-T adduct (see [6]).  相似文献   

12.
Mass spectra of the N-acetylnitramines l-acetylhexahydro-3,5-dinitro-l,3,5-triazine (TAX) and 1-acetyloctahydro-3,5,7-trinitro-l,3,5,7-tetrazocine (SEX) were recorded in electron impact (EI) and positive and negative chemical ionization (PCI and NCI) modes, and the fragmentation pathways were compared with those of other nitramines which have been well documented and characterized. Unexpectedly, for both acetylnitramines in the EI mode (and in the PCI mode) proton adducts were the only molecular ion species observed; in neither mode was there evidence for higher adducts. In contrast, for TAX in the NCI mode the [M + NO2]? adduct was the second most abundant ion (70%); relatively small amounts of the [M + NO]? adduct (2%) and the hydrogen adduct [MH]? (3%) were observed. Under identical NCI conditions no molecular ion species or adduct ions were detected for SEX; the ion of highest m/z corresponded to loss of NO2 or HNO2 from a molecular ion species. The findings of collision-induced dissociation experiments are also discussed.  相似文献   

13.
Abstract— Laser flash photolysis at 265 nm has been employed for measuring the initial hydrated electron (e-aq) and p-alanylphenoxyl radical (Tyr) in aqueous Tyr, small Tyr peptides and R Nase A. The results indicate that monophotonic photolysis not involving the fluorescent or triplet states is the principal initial process. Equivalent yields of e and Tyr were found in all cases except Tyr, where the Tyr yield was 60% higher than e-aq attributed to splitting of the phenolic bond. Computer analysis of e-aq and Tyr decays for Tyr indicates the importance of electron-radical recombination in competition with electron scavenging and bimolecular radical-radical reactions. Evidence for intramolecular electron migration has been obtained in cystinyl-bis-Tyr.  相似文献   

14.
On irradiation (300 nm) in the solid state, the title compound 8 affords tricyclo[3.3.1.13,7]decan‐2‐one (=adamantan‐2‐one; 9 ) selectively via [4+2] cycloreversion. A similar result is obtained on photolysis in solution (MeCN or acetone), also in the presence of added alkenes. On irradiation in MeOH, a solvent adduct 11 is isolated in addition to 9 . From experiments in CD3OD, it can be inferred that 11 is formed via syn‐addition of MeOH to the ground‐state (E)‐heterocycle 16 .  相似文献   

15.
利用瞬态吸收光谱技术研究了不同条件下C6H5Cl与H2O2水溶液的激光闪光光解情况, 初步考察了其瞬态物种的生长和衰减等行为. 研究表明, •OH自由基和C6H5Cl反应生成C6H5Cl-OH adduct, 其反应速率常数在近中性、酸性条件下约为(5.89±0.65)×109和(7.07±0.61)×109 L•mol-1•s-1; 其衰减则符合双分子二级反应, 速率常数2k/εl=1.1×106 s-1, 而在碱性时则为(4.34±0.51)×109 L•mol-1•s-1, 衰减呈准一级反应, 速率常数为2.11×105 s-1. 在有氧条件下, O2与C6H5Cl-OH adduct反应生成C6H5Cl-OHO2 adduct, 其反应速率常数为6.8×108 L•mol-1•s-1.  相似文献   

16.
The pH effects on the photochemical reaction of amino acids and related dipeptides with 4-nitroquinoline 1-oxide (4NQO) as a photosensitizer have been investigated by laser flash photolysis. The obtained kinetic parameters show that the electron transfer from Tryptophan (Trp), Tyrosine (Tyr) as well as dipeptides containing Trp and/or Tyr residue to triplet 4NQO (T4NQO) are efficient, but inefficient from methionine (Met) and dipeptides containing neither Trp nor Tyr. The result was supported by the calculated values of the free energy change from measured oxidation potentials for the electron transfer. It was demonstrated that Trp and Tyr residues are initial reaction sites with T4NQO, while Tyr/O? radical may be final species for Trp-Tyr dipeptide. In acidic aqueous solutions, the self-quenching rate constants of T4NQO and the rate constants of electron transfer from amino acids to T4NQO decrease with decreasing pH. In alkaline solutions, amino acids are easily oxidized by 4NQO under irradiation of laser pulse, and no transient absorption signal was observed.  相似文献   

17.
Abstract

Cationic radicals generated by anodic oxidation of organophosphorus(III) compounds are superelectrophilic radicals. This determines the pattern of their reactivity towards various substrates. The oxidation of (RO)2 PO? and (RO)2PS? gives corresponding radicals. Interaction of the mentioned radical species with ArH and olefines includes a common stage of radical adduct formation. It has been stated that the oxidation of the adduct into corresponding carbonium ion plays an important role in the following adduct conversions. This is particularly confirmed by the difference in the composition of products obtained in conditions of homogeneous (via photolysis or with stable radical cations) and heterogeneous (on the anode surface) generation of phosphorylating species.  相似文献   

18.
Negative chemical ionization mass spectrometry is used as a probe to examine reactions between hydrocarbon radicals and metal complexes in the gas phase. The methane negative chemical ionization mass spectra of 27 complexes of cobalt(II ), nickel(II ) and copper(II ) in the presence of O4, O2N2 and N4 donor atom sets are characterized by two dominant series of adduct ions of the form [M + CnH2n]? and [M + CnH2n+1]? at m/z values above the molecular ion, [M]?. Insertion of the CH radical into the ligand followed by radical/radical recombination and electron capture is proposed as the major mechanism leading to the formation of [M + CnH2n]? adduct ions. A second pathway involves ligand substitution by CnH2n+1 radicals concomitant with H elimination and electron capture. Oxidative addition at the metal followed by ionization is suggested as the principal pathway for the formation of [M + CnH2n+1]? adduct ions.  相似文献   

19.
Internal electron transfer in bovine cytochrome c oxidase was initiated by CO photolysis of the CO-bound mixed-valence form of the enzyme. Transient absorption spectroscopy was used to monitor changes in the redox states of the metal centers in the enzyme brought about by electron re-equilibration. Upon CO photodissociation, reduced high spin cytochrome a3 was generated in less than 0.1 μsec, and a portion of the reduced cytochrome a3 was reoxidized with biphasic rate constants of k1 = 1.0 × 106 s?1 and k2 = 7.8 × 104 s?1. Concomitant reduction of cytochrome a was also observed with biphasic rate constants of k1 = 1.6 × 106 s?1 and k2 = 9 × 104 s?1. The stoichiometry of cytochrome a3 oxidized to cytochrome a reduced was found to be close to 1:1. Contrary to similar studies in the literature, no reduction of CuA was observed. As a control, no transient absorption changes corresponding to electron transfer was observed in the CO-inhibited fully reduced form of the enzyme. These results indicate that there is significant electron reequilibration only between cytochrome a3 and cytochrome a upon photolysis of the CO-bound mixed-valence enzyme.  相似文献   

20.
Abstract

(1,10-Phenanthroline)bis(4-morpholinecarbodithioato-S,S′)zinc(II), [Zn(mdtc)2(1,10-phen)] ( 1 ) and (2,2′ -bipyridine)bis(4-morpholinecarbodithioato-S,S′) zinc(II), [Zn(mdtc)2(bipy)] ( 2 ) adducts were prepared, and the crystal structure of 2 is reported. The Zn-S distances in 2 are longer than those in [Zn(dtc)2](dtc = dithiocarbamate) complexes. The acceptance of an additional neutral ligand by the tetrahedral dithiocarbamato complex of zinc to form an octahedral adduct causes an increase in the Zn-S bond lengths and a consequent lowering of the S-Zn-S bite angle. The cyclic voltammetric study on the complexes shows an increase of electron density on the central metal ion in the adducts compared to Zn(mdtc)2. The use of HMDE as a working electrode in the CV studies indicate the involvement of Hg in complex formation, which is otherwise absent with a platinum working electrode. The thioureide C-N distance of 1.330 ( 4 ) Å in compound 2 is in line with the ν C?N observed at 1465 cm?1.  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