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.
Narine AA  Watson JN  Bennet AJ 《Biochemistry》2006,45(30):9319-9326
The sialidase from Micromonospora viridifaciens has been found to catalyze the hydrolysis of aryl 2-thio-alpha-D-sialosides with remarkable efficiency: the first- and second-order rate constants, kcat and kcat/Km, for the enzyme-catalyzed hydrolysis of PNP-S-NeuAc are 196 +/- 5 s(-1) and (6.7 +/- 0.7) x 10(5) M(-1) s(-1), respectively. A reagent panel of eight aryl 2-thio-alpha-D-sialosides was synthesized and used to probe the mechanism for the M. viridifaciens sialidase-catalyzed hydrolysis reaction. In the case of the wild-type enzyme, the derived Br?nsted parameters (beta(lg)) on kcat and kcat/Km are -0.83 +/- 0.11 and -1.27 +/- 0.17 for substrates with thiophenoxide leaving groups of pKa values > or = 4.5. For the general-acid mutant, D92G, the derived beta(lg) value on kcat for the same set of leaving groups is -0.82 +/- 0.12. When the conjugate acid of the departing thiophenol was < or = 4.5, the derived Br?nsted slopes for both the wild-type and the D92G mutant sialidase were close to zero. In contrast, the nucleophilic mutant, Y370G, did not display a similar break in the Br?nsted plots, and the corresponding values for beta(lg), for the three most reactive aryl 2-thiosialosides, on kcat and kcat/Km are -0.76 +/- 0.28 and -0.84 +/- 0.04, respectively. Thus, for the Y370G enzyme glycosidic C-S bond cleavage is rate-determining for both kcat and kcat/Km, whereas, for both the wild-type and D92G mutant enzymes, the presented data are consistent with a change in rate-determining step from glycosidic C-S bond cleavage for substrates in which the pKa of the conjugate acid of the leaving group is > or = 4.5, to either deglycosylation (kcat) or a conformational change that occurs prior to C-S bond cleavage (kcat/Km) for the most activated leaving groups. Thus, the enzyme-catalyzed hydrolysis of 2-thiosialosides is strongly catalyzed by the nucleophilic tyrosine residue, yet the C-S bond cleavage does not require the conserved aspartic acid residue (D92) to act as a general-acid catalyst.  相似文献   

2.
The photochemistry of 6-chloro and 6-bromopicolinate ions ( and , respectively) was investigated by product studies and ns laser flash photolysis (LFP). In deoxygenated pH 5.4 water, yields 6-hydroxypicolinic acid (70%) and a substituted pyrrole. In 2-propanol-water (1 : 1) mixture, the reaction yields, very unselectively, 6-hydroxypicolinic acid, 2-carboxypyridine, pyridine and bipyridines. Photolysis of aqueous leads to 6-hydroxypicolinic acid (78%) and hydroxybipyridines. Oxygen suppresses the photolysis of but does not affect that of . By LFP, we detected a short-lived transient at the pulse end from (lambda(max)= 305 nm, k=(2.8 +/- 0.2)x 10(5) s(-1), epsilonphi= 2200 +/- 200 dm3 mol(-1) cm(-1)). This is quenched either by oxygen or methyl acrylate and thus assigned to the triplet excited state. The triplet excited state of is detected at pH 1 only (lambda(max)= 320 nm, k > 3 x 10(7) s(-1)). The radical ion Cl2- could be successfully detected by photolysing in 2-propanol-water (1 : 1) in the presence of Cl-. Similarly, Br2- could be detected by irradiating aqueous in the presence of Br-. These results show that the photodehalogenation of is heterolytic in water and mainly homolytic in 2-propanol-water mixtures while that of is both heterolytic and homolytic in water. A mechanism in which the triplet excited state undergoes homolysis of the C-X bond and subsequent electron transfer from the carboxypyridyl radical to the halogen atom to form an ion pair may account for these observations.  相似文献   

3.
Clapp CH  McKown J  Xu H  Grandizio AM  Yang G  Fayer J 《Biochemistry》2000,39(10):2603-2611
Previous work has demonstrated that the ferric form of soybean lipoxygenase-1 will catalyze an elimination reaction on 12-iodo-cis-9-octadecenoic acid (12-IODE) to produce 9, 11-octadecadienoic acid and iodide ion. Elimination is accompanied by irreversible inactivation of the enzyme on 1 out of 10 turnovers. In the present work, 11,11-dideuterio-12-IODE (D(2)-12-IODE) was synthesized and used to demonstrate that both the elimination reaction and inactivation of the enzyme exhibit very large kinetic isotope effects. The rates with the deuterated compound are so low that the isotope effects are difficult to quantify, but they appear to be comparable to the isotope effects previously observed for the normal reaction catalyzed by lipoxygenase and much larger than can be explained by zero-point energy considerations. ESR spectroscopy was used to demonstrate that 12-IODE can reduce ferric lipoxygenase to the ferrous form, and a large isotope effect on this process was observed with D(2)-12-IODE. It is proposed that the pathway leading to reduction and inactivation by 12-IODE is initiated by homolytic cleavage of the C(11)-H bond. Elimination could be initiated either by homolytic or by heterolytic cleavage of this bond. The results suggest that very large isotope effects may be a general feature of C-H bond cleavages catalyzed by this enzyme.  相似文献   

4.
The cyclooxygenase (COX) activity of prostaglandin endoperoxide H synthases (PGHSs) converts arachidonic acid and O2 to prostaglandin G2 (PGG2). PGHS peroxidase (POX) activity reduces PGG2 to PGH2. The first step in POX catalysis is formation of an oxyferryl heme radical cation (Compound I), which undergoes intramolecular electron transfer forming Intermediate II having an oxyferryl heme and a Tyr-385 radical required for COX catalysis. PGHS POX catalyzes heterolytic cleavage of primary and secondary hydroperoxides much more readily than H2O2, but the basis for this specificity has been unresolved. Several large amino acids form a hydrophobic "dome" over part of the heme, but when these residues were mutated to alanines there was little effect on Compound I formation from H2O2 or 15-hydroperoxyeicosatetraenoic acid, a surrogate substrate for PGG2. Ab initio calculations of heterolytic bond dissociation energies of the peroxyl groups of small peroxides indicated that they are almost the same. Molecular Dynamics simulations suggest that PGG2 binds the POX site through a peroxyl-iron bond, a hydrogen bond with His-207 and van der Waals interactions involving methylene groups adjoining the carbon bearing the peroxyl group and the protoporphyrin IX. We speculate that these latter interactions, which are not possible with H2O2, are major contributors to PGHS POX specificity. The distal Gln-203 four residues removed from His-207 have been thought to be essential for Compound I formation. However, Q203V PGHS-1 and PGHS-2 mutants catalyzed heterolytic cleavage of peroxides and exhibited native COX activity. PGHSs are homodimers with each monomer having a POX site and COX site. Cross-talk occurs between the COX sites of adjoining monomers. However, no cross-talk between the POX and COX sites of monomers was detected in a PGHS-2 heterodimer comprised of a Q203R monomer having an inactive POX site and a G533A monomer with an inactive COX site.  相似文献   

5.
The photochemistry of optically pure isomers of alpha-methylbenzylamide of trans-2,3-diphenylcyclopropane-1-carboxylic acid has been examined in isotropic solution and within zeolites. The results suggest that these isomerize through cleavage of C2-C3 bond. The direct excitation in solution leads to non-equilibrating 1,3-singlet diradical intermediates whereas triplet sensitization results in equilibrating 1,3-triplet diradical intermediates. The direct excitation within NaY zeolite seems to result in equilibrating zwitterionic intermediates. Studies on the optically pure trans isomers allow one to understand the mechanism of chiral induction during the photoisomerization of mesocis-2,3-diphenylcyclopropane-1-carboxylic acid. The current study has clarified the nature of the excited states involved during the classic (R)-N-acetyl-1-naphthylethylamine sensitized isomerization of 1,2-diphenylcyclopropane.  相似文献   

6.
The photochemistry and molecular structure of 1-phenyl-4-allyl-tetrazolone (PAT) was studied by FT-IR matrix isolation spectroscopy and DFT(B3LYP)/6-311++G(d,p) calculations. The spectrum of matrix-isolated PAT monomers agrees well with the sum spectrum of three conformers predicted theoretically. UV irradiation (lambda > 235 nm) of matrix-isolated PAT induces three types of photofragmentation: (1) production of phenylazide and allyl-isocyanate, with phenylazide then losing N(2) to yield 1-aza-1,2,4,6-cycloheptatetraene; (2) formation of phenyl-isocyanate and allylazide; (3) N(2) elimination leading to formation of 1-allyl-2-phenyldiaziridin-3-one; this compound partially reacts further to form 1-allyl-1H-benzoimidazol-2(3H)-one. The observed photochemistry of the matrix-isolated PAT is distinct from the preferred photochemical fragmentation in solution, where 3,4-dihydro-3-phenylpyrimidin-2(1H)-one is produced as the primary photoproduct.  相似文献   

7.
Lignin peroxidase catalyses the H(2)O(2)-induced oxidation of 4-methoxybenzyltrimethylsilane by an electron transfer mechanism. The intermediate radical cation undergoes preferentially C(alpha)[bond]H deprotonation to give 4-methoxybenzaldehyde whereas C(alpha)[bond]Si bond cleavage is a minor fragmentation pathway and leads to 4-methoxybenzyl alcohol. Similar results are obtained in the oxidation catalysed by the water soluble model compound 5,10,15,20-tetra(N-methyl-4-pyridyl)porphyrinatoiron(III) pentachloride. Instead, in the oxidation promoted by the genuine one-electron transfer oxidant potassium dodecatungstocobalt(III)ate C(alpha)[bond]Si bond cleavage is the exclusive fragmentation process of the intermediate radical cation. It is suggested that in the enzymatic and biomimetic oxidations of 4-methoxybenzyltrimethylsilane the deprotonation of the intermediate radical cation is promoted by the reduced form [PorFe(IV)[double bond]O] of the active oxidant, which is an iron-oxo porphyrin radical cation.  相似文献   

8.
The regularities of mass spectrometric fragmentation under electron impact of new 9-(o-chlorobenzyl)-8-azahypoxantines with (N-aryl)amidocarbonylmethylthiomethyl substituents in position 2 were studied. The main fragmentation pathways are the elimination of Ar-NH+ and o-chlorobenzyl ions and cleavage of C-S bonds, characteristic of organic sulfides. During the fragmentations, some rearrangements occur, consisting in the transfer of labile hydrogen atoms from the alpha-positions to ions being eliminated. Fragmentation of 8-azapurine parts of the molecules does not prevail. Peaks of molecular ions are clearly visible in the mass spectra of all the substances studied. The English version of the paper: Russian Journal of Bioorganic Chemistry, 2004, vol. 30, no. 1; see also http://www.maik.ru.  相似文献   

9.
The (1)H-(13)C coupling constants of methyl alpha- and beta-pyranosides of D-glucose and D-galactose have been measured by one-dimensional and two-dimensional (1)H-(13)C heteronuclear zero and double quantum, phase sensitive J-HMBC spectra to determine a complete set of coupling constants ((1)J(CH), (2)J(CH), (3)J(CH), (2)J(HH), and (3)J(HH)) within the exocyclic hydroxymethyl group (CH(2)OH) for each compound. In parallel with these experimental studies, structure, energy, and potential energy surfaces of the hydroxymethyl group for these compounds were determined employing quantum mechanical calculations at the B3LYP level using the 6-311++G( * *) basis set. Values of the vicinal coupling constants involving (1)H and (13)C in the C5-C6 (omega) and C6-O6 (theta) torsion angles in the aldohexopyranoside model compounds were calculated with water as the solvent using the PCM method. To test the relationship between (3)J(CXCH) (X=C, O, S) and torsion angle C1-X (phi) around the anomeric center, the conformations of 24 derivatives of glucose and galactose, which represent sequences of atoms at the anomeric center of C-glycosides (C-C bond), O-glycosides (C-O bond), thioglycosides (C-S bond), glycosylamines (C-N bond), and glycosyl halides (C-halogen (F/Cl) bond) have been calculated. Nonlinear regression analysis of the coupling constants (1)J(C1,H1), (2)J(C5,H6R), (2)J(C5,H6S), (2)J(C6,H5), (3)J(C4,H6R), (3)J(C4,H6S), (2)J(H6R,H5), and (3)J(H5,H6R) as well as (3)J(CXCH) (X=C, O, S) on the dihedral angles omega, theta, and phi have yielded new Karplus equations. Good agreement between calculated and experimentally measured coupling constants revealed that the DFT method was able to accurately predict J-couplings in aqueous solutions.  相似文献   

10.
Fluorescence and trans-->cis photoisomerization are the main deactivation paths following excitation of trans-1-(2'-naphthyl)-2-(3'-hydroxyphenyl)ethene (trans-2,3NOH) in cyclohexane, methanol and acetonitrile. The quantum yield of both processes is wavelength dependent: this is due to the presence of conformational isomers deriving from rotation of the naphthyl group around the single bond with ethene. Addition of water to acetonitrile quenches the fluorescence (lambda(max)= 380 nm). In CH(3)CN/H(2)O (4/6, v/v) the emission spectrum displays a broad band with maximum at approximately 550 nm besides the original quenched fluorescence. This indicates that 2,3-NOH undergoes acid-base equilibration in the singlet excited state as supported by the enhancement of the fluorescence quantum yield with increasing acidity of the medium. Ground and excited state acidity constants have been determined. The main photochemical process is photo-hydration, i.e. water addition to the ethene bond. Fluorescence and photo-hydration have the same sigmoidal dependence on the acid concentration, which indicates that the undissociated form of singlet excited 2,3NOH is the photoreactive species. Laser flash photolysis experiments allowed identification of the reactive intermediates. The photophysics of trans-1-(1'-naphthyl)-2-(3'-hydroxyphenyl)ethene (trans-1,3NOH) is similar to that of 2,3NOH as regards the effect of water on fluorescence and the acid-base behaviour in the ground and first excited singlet state; the main photochemical process is trans-->cis photoisomerization together with photo-cylization to hydroxychrysene in neutral water/acetonitrile, but with lower yield compared to cyclohexane, and photo-hydration in strongly acidic medium.  相似文献   

11.
Y C Horng  D F Becker  S W Ragsdale 《Biochemistry》2001,40(43):12875-12885
Methyl-coenzyme M reductase (MCR), the key enzyme in methanogenesis, catalyzes methane formation from methyl-coenzyme M (methyl-SCoM) and N-7-mercaptoheptanoylthreonine phosphate (CoBSH). Steady-state and presteady-state kinetics have been used to test two mechanistic models that contrast in the role of CoBSH in the MCR-catalyzed reaction. In class 1 mechanisms, CoBSH is integrally involved in methane formation and in C-S (methyl-SCoM) bond cleavage. On the other hand, in class 2 mechanisms, methane is formed in the absence of CoBSH, which functions to regenerate active MCR after methane is released. Steady-state kinetic studies are most consistent with a ternary complex mechanism in which CoBSH binds before methane is formed, as found earlier [Bonacker et al. (1993) Eur. J. Biochem. 217, 587-595]. Presteady-state kinetic experiments at high MCR concentrations are complicated by the presence of tightly bound CoBSH in the purified enzyme. Chemical quench studies in which (14)CH(3)-SCoM is rapidly reacted with active MCRred1 in the presence versus the absence of added CoBSH indicate that CoBSH is required for a single-turnover of methyl-SCoM to methane. Similar single turnover studies using a CoBSH analogue leads to the same conclusion. The results are consistent with class 1 mechanisms in which CoBSH is integrally involved in methane formation and in C-S (methyl-SCoM) bond cleavage and are inconsistent with class 2 mechanisms in which CoBSH binds after methane is formed. These are the first reported pre-steady-state kinetic studies of MCR.  相似文献   

12.
The copper(ii) complexes of two salicylaldehyde semicarbazones, HOC(6)H(4)CH[double bond, length as m-dash]N-NHCONR(2) [H(2)Bnz(2) (R = CH(2)Ph) and H(2)Bu(2) (R = Bu)], were evaluated for their DNA binding and cleavage properties by spectrophotometric DNA titration, ethidium bromide displacement assay and electrophoretic mobility shift assay. Results showed that the Cu(ii) complexes can bind to DNA via a partial intercalation mode with binding constants of 1.1 × 10(4) and 9.5 × 10(3) M(-1) for [Cu(HBnz(2))Cl] and [Cu(HBu(2))Cl], respectively. These complexes also cleave DNA in the presence of ascorbic acid, most likely through hydroxyl radicals that are generated via the reduction of a Cu(ii) to a Cu(i) species. The complexes show similar DNA cleavage activity, which is reflected in the similarity of their frontier molecular orbital energies calculated by density functional theory. These results are discussed in relation to the anticancer properties of the complexes.  相似文献   

13.
Odd numbered 1-alkenes, such as 1-pentadecene or 1,8,11,14-heptadecatetraene are formed from palmitic or linolenic acid by fragmentative decarboxylation. Incubation studies with germinating safflower (Carthamus tinctorius) and (2R,3R)-12-phenyl[2,3-2H2]dodecanoic acid, (2S,3S)-12-phenyl[2,3-2H2]dodecanoic acid, (2R)-12-phenyl[2-2H]dodecanoic acid and (2S)-12-phenyl[2-2H]dodecanoic acid instead of the natural alpha-linolenic acid precursor revealed the fragmentation to be an overall anti elimination of the 3-pro(S) hydrogen and the carboxyl group (anti-periplanar transition state geometry). Externally offered 3-hydroxy acids are not fragmented to 1-alkenes. The most probable mechanistic alternatives are in agreement with abstraction of the 3-pro(S) hydrogen as a radical followed by electron transfer and fragmentation, or transient insertion of oxygen into the 3-pro(S) C-H bond and subsequent fragmentation into an 1-alkene and CO2 after appropriate activation. The mechanism seems to be of general importance for the biosynthesis of vinylic substructures of natural products from oxygenated precursors.  相似文献   

14.
A wide range of experimental data is evaluated in support of the hypothesis that the reductive deamination of amino acids and peptides by eaq- and the oxidative dephosphorylation of glycol phosphates by OH in oxygen free solution are related in terms of a common elimination reaction involving free-radical intermediates of the same genre, that is, XCH(R)C(OH)R----HX + CH(R)COR where X = NH2, RCONH, and H2PO4 respectively. The proposed reaction model unifies for the first time the basic free-radical chemistry of N-C and PO-C bond cleavage in the radiolysis of peptides and glycol phosphates including proteins and nucleic acids in oxygen-free solution.  相似文献   

15.
Single-stranded oligoribonucleotides containing UA and CA phosphodiester bonds can be hydrolyzed specifically under non-enzymatic conditions in the presence of spermidine, a biogenic amine found in a wide variety of organisms. In the present study, the rate of oligonucleotide and tRNA(i)(Met)hydrolysis was measured in the presence of spermidine and other biogenic amines. It was found that spermine [H(3)N(+)(CH(2))(3)(+)NH(2)(CH(2))(4)(+)NH(2)(CH(2))(3)(+)NH(3)] and putrescine [H(3)N(+)(CH(2))(4)(+)NH(3)] can replace spermidine [H(3)N(+)-(CH(2))(4)(+)NH(2)(CH(2))(3)(+)NH(3)] to induce the hydrolysis. For all three polyamines, a bell-shaped cleavage rate versus concentration relationship was observed. The maximum rate of hydrolysis was achieved at 0.1, 1.0 and 10 mM spermine, spermidine and putrescine, respectively. Moreover, we found that the hydrolysis requires at least two linked amino groups since two aminoalcohols, 2-aminoethanol and 3-aminopropanol, were not able to induce the cleavage of the phospho-diester bond. The optimal cleavage rate of the oligo-ribonucleotides was observed when amino groups were separated by tri- or tetramethylene linkers. The methylation of the amino groups reduced the ability of diamines to induce oligoribonucleotide hydrolysis. Non-enzymatic cleavage of tRNA(i)(Met)from Lupinus luteus and tRNA(i)(Met)from Escherichia coli demonstrate that both RNAs hydrolyze as expected from principles derived from oligoribonucleotide models.  相似文献   

16.
Rhodococcus erythropolis IGTS8 that possesses dibenzothiophene sulfone monooxygenase mutated at residue 345 (Q345A), can degrade octyl sulfide on which the wild strain cannot grow. Residue 345 and the neighbouring residues were changed by site-directed mutagenesis. Only DszA changed at residue 345 gave an altered C-S bond cleavage pattern of 3-methyl DBT sulfone. This residue is therefore involved in C-S bond cleavage specifically for alkylated DBT sulfone.  相似文献   

17.
Borowski T  Bassan A  Siegbahn PE 《Biochemistry》2004,43(38):12331-12342
Density functional calculations using the B3LYP functional has been used to study the reaction mechanism of 4-hydroxyphenylpyruvate dioxygenase. The first part of the catalytic reaction, dioxygen activation, is found to have the same mechanism as in alpha-ketoglutarate-dependent enzymes; the ternary enzyme-substrate-dioxygen complex is first decarboxylated to the iron(II)-peracid intermediate, followed by heterolytic cleavage of the O-O bond yielding an iron(IV)-oxo species. This highly reactive intermediate attacks the aromatic ring at the C1 position and forms a radical sigma complex, which can either form an arene oxide or undergo a C1-C2 side-chain migration. The arene oxide is found to have no catalytic relevance. The side-chain migration is a two-step process; the carbon-carbon bond cleavage first affords a biradical intermediate, followed by a decay of this species forming the new C-C bond. The ketone intermediate formed by a 1,2 shift of an acetic acid group rearomatizes either at the active site of the enzyme or in solution. The hypothetical oxidation of the aromatic ring at the C2 position was also studied to shed light on the 4-HPPD product specificity. In addition, the benzylic hydroxylation reaction, catalyzed by 4-hydroxymandelate synthase, was also studied. The results are in good agreement with the experimental findings.  相似文献   

18.
S Hammerum  C Djerassi 《Steroids》1975,25(6):817-826
The mass spectral fragmentation of a number of 17alpha-hydroxy-, 17alpha-acetoxy-, and 17alpha-methoxyprogesterones have been examined. Unlike other steroidal delta4-3-ketones, fragmentation reactions associated with the alpha,beta-unsaturated ketonic function are not particularly significant; rather, abundant ions are formed by decomposition processes occurring in and around ring D. Reactions of diagnostic significance include complete or partial loss of ring D, and elimination of the C-17 side chain (CH3CO), followed by loss of the C-17 oxygen function together with a hydrogen atom (H2O, CH3COOH, CH3OH).  相似文献   

19.
The p-peroxyquinol derived from butylated hydroxytoluene, 2,6-di-t-butyl-4-hydroperoxy-4-methyl-2,5-cyclohexadienone, was degraded by the ferric form of rat liver cytochrome P-450, and the resulting products and their mechanisms of formation were investigated. Quinoxy radical BO. from homolysis of the O-O bond reacted by competing pathways; beta-scission yielded 2,6-di-t-butyl-p-benzoquinone, and rearrangement with ring-expansion produced an oxacycloheptadienone free radical (X(.)). This rearranged radical was stabilized by the captodative effect that facilitated competitive interactions with the P-450 iron-oxo complexes formed during O-O bond scission. Approximately 15% of X(.) was captured by oxygen rebound with a hydroxyl radical from the P-450 complex (FeOH)3+ to form a hemiketal, that led to the ring-contracted product 2,5-di-t-butyl-5-(2'-oxopropyl)-4-oxa-2-cyclopentenone by spontaneous rearrangement. The major fraction of X(.), however, underwent electron transfer oxidation to form the corresponding cation. Hydration of this cation produced the ring-contracted product, and proton elimination (or, alternatively, direct H(.) removal from X(.) led to the product 2,7-di-t-butyl-4-methylene-5-oxacyclohepta-2,6-dienone. The findings indicate that cytochrome P-450 intermediate complexes are mainly responsible for oxidation of X(.). The results complement our previous study with 2,6-di-t-butyl-4-hydroperoxy-4-methyl-2,5-cyclohexadienone (Thompson, J. A., and Wand, M. D. (1985) J. Biol. Chem. 260, 10637-10644), demonstrating competitive heterolytic and homolytic mechanisms of O-O bond cleavage, and competitive rebound and oxidation processes when a substrate-derived radical interacts with P-450 complexes.  相似文献   

20.
We present measurements of low-energy (<18 eV) electron-stimulated desorption of anions from acetamide (CH(3)CONH(2)) and dimethyl disulfide [DMDS: (CH(3)S)(2)] films. Electron irradiation of physisorbed CH(3)CONH(2) produces H(-), CH(3)(-) and O(-) anions, whereas the H(-), CH(2)(-), CH(3)(-), S(-), SH(-) and SCH(3)(-) anions are observed to desorb from the DMDS film. Below 12 eV, the dependence of the anion yields on the incident electron energy exhibits structures that indicate that a resonant process (i.e. dissociative electron attachment) is responsible for molecular fragmentation. Within the range of 1-18 eV, it is found that (1.7 and 1.4) x 10(7) H(-) ions/incident electron and (7.8 x 10(-11) and 4.3 x 10(-8)) of the other ions/incident electron are desorbed from acetamide and DMDS films, respectively. These results suggest that, within proteins, the disulfide bond is more sensitive to low-energy electron attack than the peptide bond. In biological cells, some proteins interact closely with nucleic acid. Therefore, the observed fragments, when produced from secondary low-energy electrons generated by high-energy radiation, not only may denature proteins, but may also induce reactions with the nearby nucleic acid and damage DNA.  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