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 46 毫秒
1.
The impact of low-energy (1-30 eV) electrons on self-assembled monolayers of heterogeneous oligonucleotides chemisorbed on a gold surface has been investigated by mass spectrometry of desorbed neutral species in an attempt to understand the consequences of secondary electron damage in a short sequence of a DNA single strand. We demonstrate that the most intense observable neutral species (CN, OCN and/or H(2)NCN) desorbed from Cy(6)-Th(3) and Cy(6)-(BrdU)(3) oligos are related to primary fragmentation of the bases induced by electron impact. The dependence of the neutral species desorption on electron energy shows typical signatures of dissociative electron attachment initiated by the formation of shape- and core-excited resonances (i.e. single-electron and two-electron- one-hole transitory anions, respectively). Substitution of dTh by BrdU increases the production of neutral fragments by as much as a factor of about 3 for the entire electron energy range. When the distribution of secondary electrons along radiation tracks in H(2)O is taken into account, we show that the probability for electron damage to heterogeneous oligonucleotides is enhanced by a factor of 2.5-3 for electron energies below 20 eV for both sensitized and unsensitized strands.  相似文献   

2.
We present the results of experiments on anion desorption from the physisorbed DNA bases adenine, thymine, guanine and cytosine induced by the impact of low-energy (5-40 eV) electrons. Electron bombardment of DNA base films induces ring fragmentation and desorption of H(-), O(-), OH(-), CN(-), OCN(- ) and CH(2)(-) anions through either single or complex multibond dissociation. We designate the variation of the yield of an anion with electron energy as the yield function. Below 15 eV incident electron energy, bond cleavage is controlled mainly by dissociative electron attachment. Above 15 eV, the portion of a yield function that increases linearly is attributed to nonresonant processes, such as dipolar dissociation. A resonant structure is superimposed on this signal around 20 eV in the anion yield functions. This structure implicates dissociative electron attachment and/or resonant decay of the transient anion into the dipolar dissociation channel, with a minimal contribution from multiple inelastic electron scattering. The yields of all desorbing anions clearly show that electron resonances contribute to the damage of all DNA bases bombarded with 5-40 eV electrons. Comparison of the ion yields indicates that adenine is the least sensitive base to slow electron attack. Electron-irradiated guanine films exhibit the largest yields of desorbed anions.  相似文献   

3.
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   

4.
Within the last two decades, a number of experimental techniques have been developed to measure mean free paths and absolute and effective cross-sections for various processes related to the interaction of low-energy electrons with condensed matter. In all of the experiments, a monochromatic electron beam impinges on a thin multilayer film composed of atoms and/or molecules condensed on a metal or semiconductor substrate held at cryogenic temperatures in an ultra-high-vacuum system. Depending on the apparatus, cross-sections are obtained from low-energy electron transmission (LEET), high-resolution electron energy loss (HREEL), x-ray photoelectron (XPS) spectroscopy, electron-stimulated desorption (ESD) of neutral and ions, or a combination of these techniques. Quasi-elastic and inelastic mean free paths have been extracted from LEET data. This method has also served to generate absolute cross-sections for electron trapping and fragment production from the dissociation of transient molecular anions. In amorphous ice, a complete set of absolute cross-sections for all inelastic losses by 1–20 eV electrons has been obtained from HREEL data. Effective cross-sections for neutral and ionic radical formation were generated by desorption and XPS experiments. These various methods are briefly described in this article, and the corresponding cross-sections in the range 0–20 eV summarized. Received: 10 September 1998 / Accepted: 22 October 1998  相似文献   

5.
The ability of low-energy electrons to induce single- and double-strand breaks in DNA has recently been demonstrated. Here we show the propensity of 3-30 eV electrons to initiate base sequence-dependent damage to a short single DNA strand. Solid monolayer films of homogeneous thymidine (T(9)) and deoxycytidine (dCy(9)) and heterogeneous oligomers (T(6)dCy(3)) are bombarded with 1-30 eV electrons in an ultrahigh-vacuum system. CN, OCN and/or H(2)NCN are detected by a mass spectrometer as the most intense neutral fragments desorbing in vacuum. A weaker signal of CH(3)CCO is also detected, but only from oligonucleotides containing thymine. Below 17 eV, the energy dependence of the yields of CN, OCN and CH(3)CCO exhibits resonance-like structures, attributed to dissociative electron attachment (DEA). Above 17 eV, the monotonic increase in the fragment yields indicates that nonresonant processes (i.e. dipolar dissociation) control the fragmentation of these molecules. Within the energy range investigated, comparison of the magnitude of the total fragment yields produced by electron attack on dCy(9), T(6)-dCy(3) and T(9) suggests the following order in the sensitivity of single-strand DNA: dCy(9) > T(6)-dCy(3) > T(9). At 12 eV, the total fragment yields are found to be 5.8, 5.0 and 3.9 x 10(-3) fragment/electron, respectively. From the yields obtained with the two homo-oligonucleotides, we differentiate between contributions arising from the chemical nature of the base and the effect of environment (i.e. the sequence) when a thymidine unit in T(9) is replaced by dCy. The base sequence-dependent damage is found to vary with incident electron energy. These results reinforce the idea that genomic sensitivity to ionizing radiation depends on local genetic information. Furthermore, they underscore the possible role of low-energy electrons in the pathways responsible for the induction of specific genomic lesions.  相似文献   

6.
We report measurements of the desorption of anions from thin condensed films of tryptophan (Trp), histidine (His) and proline (Pro) stimulated by 5-35 eV electron impact. H-, O-, OH- and CN- desorb from Trp, His and Pro, whereas CH2- is observed only from Pro fragmentation. Below 12 eV, the anion yield functions exhibit resonant structures indicative of dissociative electron attachment. For all three amino acids, this process is likely to be initiated by the resonant capture of the incident electron at the NH3(+)-CH-.....-COO- and/or NH2-CH-.....-COOH group of the molecule. Temporary electron attachment to the ring leads to anion desorption only for tryptophan and proline. The energy-averaged yields measured at the detector of the mass spectrometer are (4.9, 0.3 and 54.0) x 10(-8) H-/incident electron and (3.4, 2.9, 1.8) x 10(-11) O-/incident electron, respectively, from Trp, His and Pro dissociation. Fragmentation of amino acids is found to be as intense as that of the nucleic acid bases. These results are discussed within the context of radiobiological damage induced by secondary electrons.  相似文献   

7.
Positive ion desorption from thin films of DNA components, 2-deoxy-d-ribose, thymine, thymidine (dThd), and thymidine 5'-monophosphate (dTMP) was investigated in the oxygen K- shell edge excitation region using synchrotron ultrasoft X rays (538 eV). A large number of molecular fragments, H(+), CH(x)(+), C(2)H(x)(+), CO(+), CH(x)O(+), C(3)H(x)(+), C(2)H(x)O(+) and C(3)H(x)O(+) (x = 1, 2 and 3), were observed as desorbed ions from 2-deoxy-d-ribose. Some of these ions are related to simultaneous bond scission at particular C-C and C-O (or C-C) bonds in the furanose ring structure in the 2-deoxy-d-ribose molecule, indicating that the impact of photons on the oxygen atom and the impact of ejected secondary electrons (e.g. Auger electrons) cause an intense destruction of the furanose ring structure. In thymine thin films, H(+), CH(x)(+), CO(+), CH(x)O(+), C(2)H(x)N(+) and CH(x)NO(+) (x = 1, 2 and 3) fragments were observed. The yields of these ions were smaller than the yields from 2-deoxy-d-ribose. The desorption of CH(3)(+) from thymine might induce a molecular conversion from thymine to uracil. The mass patterns of dThd and dTMP, and especially that of dTMP, were similar to that of 2-deoxy-d-ribose, indicating that a number of ions were generated at the sugar site, even in the nucleotide molecule. It is therefore predicted that the sugar moiety is more fragile than the thymine base.  相似文献   

8.
Monte Carlo track structure simulations were performed to investigate the effect of multiple ionization of water on the primary (or "escape") (at approximately 10(-6) s) yield of hydrogen peroxide (G(H2O2)) produced in the radiolysis of deaerated 0.4 M H2SO4 solutions by 12C6+ and 20Ne9+ ions at high linear energy transfer (LET) up to approximately 900 keV/microm. It was found that, upon incorporating the mechanisms of double, triple and quadruple ionizations of water in the calculations, a quantitative agreement between theory and experiment can be obtained. The curve for G(H2O2) as a function of LET reaches a well-defined maximum of approximately 1.4 molecules/100 eV at approximately 180-200 keV/microm, in very good accord with the available experimental data. Our results also show that, for the highest LET values considered in this study, the H2O2 escape yields obtained in 0.4 M sulfuric acid solutions are about 45% greater in magnitude than those found in neutral water. Contrary to a recent assumption suggesting that the limiting value of G(H2O2) at infinite LET should be approximately 1 molecule/100 eV, somewhat similar for neutral and acidic water, our simulations show a clear decrease in the primary H2O2 yields with increasing LET at high LET, indicating that the question of the limiting value of G(H2O2) at very high LET for both neutral and acidic liquid water is still open.  相似文献   

9.
Formic acid solutions of 1, 10, 100, and 1000 mM have been irradiated with 4He ions of 5 to 25 MeV, and the production of OH radicals has been determined by measuring the yield of CO2. The differential OH radical yields were obtained from the observed energy dependencies; with 25 MeV 4He ions they range from 1.91 to 3.48 molecules/100 eV for formic acid concentrations of 1 to 1000 mM, respectively. The OH radical yields decrease with decreasing particle energy, and at the maximum LET (230 eV/nm) they range from 0.30 at 1 mM to 0.82 molecules/100 eV at 1000 mM. These values are only 15 to 20% of that found with fast electrons. The OH radical yields are relatively more dependent on formic acid concentration at higher 4He ion energies. The average time dependencies of the OH radical from 7.7 ns to 7.7 microseconds were estimated from the formic acid concentration dependencies at various 4He energies. In terms of absolute yields, there is a considerable variation in the yields of OH radicals with time at the highest energies, but at the maximum LET the OH radical yields are nearly invariant with time after about 10 ns.  相似文献   

10.
Hot plasmas can be generated by fast and intense laser pulses ablating solids placed in vacuum. A Nd:Yag laser operating at the fundamental and second harmonics with 9-ns pulses (maximum energy of 900 mJ) focused on metallic surfaces produces high ablation yields of the order of μg/pulse and dense plasma that expands adiabatically at supersonic velocity along the normal to the target surface. The plasma emits neutral and charged particles. Charge states up to 10+ have been measured in heavy elements ablated with intensities of the order of 1010 W/cm2. The ion temperature of the plasma is evaluated from the ion energy distributions measured with an ion energy analyzer. The electron temperature is measured through Faraday cups placed at the end of long drift tubes by using time-of-flight technique. The neutral temperature is measured with a special mass quadrupole spectrometer placed along the normal to the target surface. The plasma temperature increases with the laser pulse intensity. The ion temperature reaches values of the order of 400 eV, the electron temperature is of the order of 1 keV for hot electrons and 0.1 eV for thermal electrons, and the neutral temperature is of the order of 200 eV. The experimental apparatus, the diagnostic techniques, and the procedures for the plasma temperature characterization will be presented and discussed in detail. Published in Russian in Fizika Plazmy, 2006, Vol. 32, No. 6, pp. 558–564. The text was submitted by the authors in English.  相似文献   

11.
The radiation-induced process of strand breaks on pBR322 plasmid DNA in aqueous solution for different energy electrons was studied by Monte Carlo simulation. Assumptions of induction mechanisms of single- and double-strand breaks (SSBs and DSBs) used in the simulation are that SSB is induced by OH or H reaction with DNA and that DSB is induced by two SSBs on the opposite strands within 10 bp. Dose-response relationships of SSBs and DSBs were demonstrated for monoenergetic electrons of 100 eV, 10 keV, 1 keV and 1 MeV, and the yields of SSB and DSB were calculated. The dose-response relationships of SSBs and DSBs can be fitted by linear and linear-quadratic functions, respectively. The ratio of quadratic to linear components of DSB induction changes due to the electron energy. A high contribution of the linear component is observed for 1 keV electrons in the dose range below 160 Gy. The yields of SSBs and DSBs for all examined electron energies lie well within the experimental data when the probability of strand-break induction by OH and H is assumed to be around 0.1-0.2. The yield of SSBs has a minimum at 1 keV, while the yield of DSBs has a maximum at 1 keV in the examined energies. The strand breaks are formed most densely for 1 keV electrons.  相似文献   

12.
Monte Carlo simulations of the radiolysis of neutral liquid water and 0.4 M H(2)SO(4) aqueous solutions at ambient temperature are used to calculate the variations of the primary radical and molecular yields (at 10(-6)s) as a function of linear energy transfer (LET) in the range approximately 0.3 to 6.5 keV/micrometer. The early energy deposition is approximated by considering short (approximately 20-100 micrometer) high-energy (approximately 300-6.6 MeV) proton track segments, over which the LET remains essentially constant. The subsequent nonhomogeneous chemical evolution of the reactive species formed in these tracks is simulated by using the independent reaction times approximation, which has previously been used successfully to model the radiolysis of water under various conditions. The results obtained are in good general agreement with available experimental data over the whole LET range studied. After normalization of our computed yields relative to the standard radical and molecular yields for (60)Co gamma radiation (average LET approximately 0.3 keV/micrometer), we obtain empirical relationships of the primary radiolytic yields as a function of LET over the LET range studied. Such relationships are of practical interest since they allow us to predict a priori values of the radical and molecular yields for any radiation from the knowledge of the average LET of this radiation only. As an application, we determine the corresponding yields for the case of (137)Cs gamma radiation. For this purpose, we use the value of approximately 0.91 keV/micrometer for the average LET of (137)Cs gamma rays, chosen so that our calculated yield G(Fe(3+)) for ferrous-ion oxidation in air-saturated 0.4 M sulfuric acid reproduces the value of 15.3 molecules/100 eV for this radiation recommended by the International Commission on Radiation Units and Measurements. The uncertainty range on those primary radical and molecular yields are also determined knowing the experimental error (approximately 2%) for the measured G(Fe(3+)) value. The following values (expressed in molecules/100 eV) are obtained: (1) for neutral water: G(e(-)(aq)) = 2.50 +/- 0.16, G(H(.)) = 0.621 +/- 0.019, G(H(2)) = 0.474 +/- 0.025, G((.)OH) = 2.67 +/- 0.14, G(H(2)O(2)) = 0.713 +/- 0.031, and G(-H(2)O) = 4.08 +/- 0.22; and (2) for 0.4 M H(2)SO(4) aqueous solutions: G(H(.)) = 3.61 +/- 0.09, G(H(2)) = 0.420 +/- 0.019, G((.)OH) = 2.78 +/- 0.12, G(H(2)O(2)) = 0.839 +/- 0.037, and G(-H(2)O) = 4.46 +/- 0.16. These computed values are found to differ from the standard yields for (60)Co gamma rays by up to approximately 6%.  相似文献   

13.
The yields in molecules per 100 eV for active-site and sulphydryl loss from glyceraldehyde-3-phosphate dehydrogenase have been determined in nitrous-oxide-saturated, aerated and argon-saturated solutions. Molecular hydrogen peroxide produces a sulphenic acid product, which can be repaired by post-irradiation treatment with dithiothreitol. Comparison of the yields under various conditions showed that in aerated solutions both .OH and .O2-radicals inactivated the enzyme with an efficiency of about 26 per cent. However, the efficiency of .OH in air-free solutions was less, and inactivation by .H and eaq- did not appear to be appreciable. There is a correlation between SH loss and loss of active sites.  相似文献   

14.
Endonuclease (Endo) IV encoded by denB of bacteriophage T4 is an enzyme that cleaves single-stranded (ss) DNA in a dC-specific manner. Previously we have demonstrated that a dTdCdA is most preferable for Endo IV when an oligonucleotide substrate having a single dC residue is used. Here we demonstrate that Endo IV cleaves ssDNAs exclusively at the 5′-proximal dC where a sequence comprises dC residues both at the 5′ proximal and 3′ proximal positions (a dCs tract-dependent cleavage). The dCs tract-dependent cleavage is efficient and occurs when a dCs tract has at least 6 bases. Some dCs tracts larger than 6 bases behave as that of 6 bases (an extended dCs tract), while some others do not. One decameric dCs tract was shown to be cleavable in a dCs tract-dependent manner, but that with 13 dCs was not. The dCs tract-dependent cleavage is enhanced by the presence of a third dC residue at least for a 6 or 7 dCs tract. In contrast to the dCs tract-dependent cleavage, a dCs tract-independent one is generally inefficient and if two modes are possible for a substrate DNA, a dCs tract-dependent mode prevails. A model for the dCs tract-dependent cleavage is proposed.  相似文献   

15.
When titanium(III) citrate was used as electron donor for the reduction of methyl coenzyme M by the methyl coenzyme M methylreductase system of Methanobacterium thermoautotrophicum delta H, component A1 was no longer required. The simpler system thus obtained required components A2, A3, and C as well as catalytic amounts of ATP, vitamin B12, and the disulfide of 7-mercaptoheptanoylthreonine phosphate in addition to titanium(III) citrate. This three component enzyme system also could produce CH4 when stoichiometric amounts of 7-mercaptoheptanoylthreonine phosphate were used as a source of electrons under an H2 atmosphere. When 7-mercaptoheptanoylthreonine phosphate or H2 was used alone no CH4 was produced, indicating a dual requirement for reducing equivalents: one to activate the methylreductase system and the other to reduce methyl coenzyme M. This is the first evidence that the activation of methyl coenzyme M methylreductase is a reductive process.  相似文献   

16.
Proton and phosphorus nmr have been used to investigate the double-helical structures of polyriboadenylic acid [poly(A)] formed in acidic solutions (pH < 6). The results obtained at low pH (~4.5) are consistent with the model for the acid poly(A) double helix proposed by Rich [Rich, A., Davies, D. R., Crick, F. H. C. & Watson, J. D. (1961) J. Mol. Biol. 3 , 71–86]. Other models that have been proposed are inconsistent with the nmr data. The nmr measurements have also been used to examine the conformation of poly(A) helix in the half-protonated state. Although the base-stacking arrangement of this state is similar to that observed in the more extensively protonated low-pH state, the phosphate backbone conformation is different from that found in either the neutral or low-pH structures.  相似文献   

17.
Polyadenylic acid (poly A) was studied under various conditions using both DC polarography and phase sensitive AC polarography and by measuring the time-course of the current during the lifetime of a single drop of the dropping mercury electrode. Under certain conditions the current at potentials of the limiting portion of the DC polarographic wave does not reach its limiting value and in extreme situations peak-shaped curves are observed. This phenomenon is explained in terms of desorption and repulsion from the electrode of neutral poly A due to its polyanionic character. Consequently, the suppression of the current can be enhanced by increasing negative potential of the electrode and by exposing the negative charges of phosphate groups, e.g., by increasing pH and temperature and by decreasing ionic strength and buffer capacity; vice versa, the current suppression can be at least partially eliminated by reversing these conditions. Polyamines which seem to shield the phosphate groups through specific interactions are very effective in eliminating the current suppression. The effectiveness of a polyamine is determined by its chain length and by the density of its amino groups and the geometry of their distribution.  相似文献   

18.
The radiolysis of aqueous chloride solutions has been investigated using pulse and steady-state methods. We have found a correlation between the yields of Cl2- and HOCl formed in pulse-irradiated N2O-saturated solutions. The yields increased with the increasing concentrations of Cl- and phosphate. Phosphate enhanced the yield of Cl2- in neutral solutions because of a proton transfer from H2PO4- to HOCl- with a rate constant of (2.6 +/- 0.5) x 10(8) M-1s-1. HOCl could not be detected in pulse-irradiated He or air-saturated, phosphate-buffered saline (PBS) solutions or in gamma-irradiated N2O, He, or air-saturated PBS solutions. The results are discussed in light of previously suggested mechanisms for the formation and decay of Cl2-. Pulse-irradiated N2O-saturated PBS solutions have a lethal effect on Escherichia coli cells, which is proportional to the amount of HOCl in the solutions. Gamma-irradiation of cells in N2O-saturated PBS solution also raises the radiosensitivity of the cells, although HOCl does not accumulate in this system. The effects of the radiation-induced toxic products on E. coli cells are similar to the effects of NaOCl. The cell membrane is probably the site of physiological injury induced by the radiation products.  相似文献   

19.
The aggregation of meso-tetra(4-sulfonatophenyl)porphyrin (H(2)TPPS(4-)) in phosphate solutions was investigated as a function of pH, concentration, time, ionic strength, and solution preparation (either from dilution of a freshly prepared 2 mM stock or by direct preparation of μM solution concentrations) using a combination of complementary analytical techniques. UV-vis and fluorescence spectroscopy indicated the formation of staggered, side-by-side (J-type) assemblies. Their size and self-associative behavior were determined using analytical ultracentrifugation and small-angle X-ray scattering. Our results indicate that in neutral and basic solutions of H(2)TPPS(4-), porphyrin dimers and trimers are formed at micromolar concentrations and in the absence of NaCl to screen any ionic interactions. At these low concentrations and pH 4, the protonated H(4)TPPS(2-) species self-assembles, leading to the formation of particularly stable aggregates bearing 25 ± 3 macrocycles. At higher concentrations, these structures further organize or reorganize into tubular, rod-like shapes of various lengths, which were imaged by cryogenic and freeze-fracture transmission electron microscopy. Micron-scale fibrillar aggregates were obtained even at micromolar concentrations at pH 4 when prepared from dilution of a 2 mM stock solution, upon addition of NaCl, or both.  相似文献   

20.
The chemistry of water and aqueous solutions is very different after irradiation with (3)H beta particles and high-energy electrons or (60)Co gamma rays. The greater the linear energy transfer (LET) of the medium for (3)H beta particles compared to high-energy electrons or (60)Co gamma rays leads to an increased local concentration of reactants. There is an increased amount of intratrack chemistry, which reduces the escape yield of and OH by about 50%, but increases the yield of H(2) by about 50% and of H(2)O(2) by about 35%. Analysis of stochastic-diffusion kinetic calculations employing simulated track structures reveals that the yield of H(2) produced by diffusion-kinetic processes increases significantly for (3)H beta particles compared to (60)Co gamma radiation, while production of H(2) by sub-picosecond processes is essentially the same. In both (3)H beta-particle and (60)Co gamma radiolysis, the reactions + and are equally important in the production of H(2). In the former case, each reaction has a yield of approximately 0.18, and in the latter a yield of approximately 0.08. In neutral water, the reaction (H + H) is negligible. The yield of Fe(III) in (3)H beta-particle radiolysis of the Fricke dosimeter is much smaller than in radiolysis with more energetic electrons. Simulations show that this change is primarily due to the reduced escape yield of H, formed from the scavenging of by the bulk H(3)O(+) of the acid. The chemical differences observed in experiments, and in calculations, reflect the underlying structure of the electron tracks: Examination of the track structure simulations demonstrates that primary events are considerably more well-separated in high-energy electron tracks compared to (3)H beta-particle tracks.  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