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.
Radiation-induced free radical formation in single crystals of guanine hydrochloride dihydrate has been studied at temperatures between 20 and 300 K using ESR and ENDOR spectroscopy. At low temperatures three radical species are trapped. Two of these are the C8 H-addition radical R1 previously analysed by Alexander and Gordy (1967) and the O6-protonated anion radical R2. The third species (R4) remains unidentified. Upon annealing at 280 K for an extended period the protonated anion R2 transforms into a new radical R3 which exhibit a well-defined hyperfine pattern but still could not be identified unambiguously. Also radical R4 probably transforms into a new radical (R5) upon such treatment. One proton coupling due to R5 was detected. A scheme of radical reactions incorporating these five radicals is proposed. This scheme also suggests that differences in radical formation between the monohydrate and dihydrate crystals of guanine hydrochloride depends upon differences in the hydrogen bonding network.  相似文献   

2.
Single crystals of guanine hydrobromide monohydrate, in which the guanine base is protonated at N7, were X-irradiated at 8 and 65 K. Using K-band ESR, ENDOR, and field-swept-ENDOR (FSE) techniques, the crystals were studied between 8 K and room temperature. There was evidence for five different radicals, RI-RV, immediately following irradiation at 8 or 65 K. RI was identified as the O6-protonated anion. It decayed near room temperature with no detectable successor. RII was identified as the N7-deprotonated cation, and decayed near 130 K. RIII is thought to be a ring-opened product formed by C8-N9 bond rupture; upon warming, it decayed at 150 K. RIV is the well-known C8 H-addition radical. These four radicals have been observed previously in the hydrochloride salt of guanine monohydrate. RV is novel, however, with magnetic characteristics consistent with those of the product formed by net OH addition to C5 of the unsaturated C4-C5 bond. It is characterized by four alpha-proton couplings indicating pi-electron spin as follows: 13% at C8; 11% at N7; and 12% at N10. RV decayed between 240 and 255 K with no detectable successor. Upon further warming, very weak resonance lines due to additional, unidentified radicals were observed. A comparison of these results with those obtained from other systems containing N7-protonated guanine bases demonstrates the effect of the environment on the primary radical formation.  相似文献   

3.
Single crystals of the four aromatic bioamine salts phenylethylamine hydrochloride, tyramine hydrochloride, tryptamine hydrochloride, and histamine dihydrochloride were grown in various states of deuteration. Free radicals were produced by exposure to X-rays between 77 and 300 K and investigated by electron spin resonance spectroscopy. Dissociation of atomic hydrogen from C beta of the aliphatic chain occurs in all compounds studied except tryptamine. However deamination as usually present in the analogous amino acids is not found. The C beta-radical is characterized by an anisotropic H alpha-splitting and two isotropic H beta-splittings. The latter splittings depend strongly on temperature in tyramine. In comparison to the analogous amino acids, radical formation in the aromatic residues is favoured. Among the seven different aromatic radicals found only one is identified in histamine but two in each of the three other bioamines. Two of these are characterized by hydrogen dissociation which occurs in phenylethylamine and tyramine. One hydrogen addition radical is found in each of the three compounds phenylethylamine, tyramine and histamine. In tryptamine two different addition radicals are detected. One of the two products can be converted into the other by visible light. The reverse process is induced by heat, thus permitting the switching of the radical site reversibly between two different structures.  相似文献   

4.
The effect of counterion was evaluated on the photodegradation behavior of six prazosin salts, viz., prazosin hydrochloride anhydrous, prazosin hydrochloride polyhydrate, prazosin tosylate anhydrous, prazosin tosylate monohydrate, prazosin oxalate dihydrate, and prazosin camsylate anhydrous. The salts were subjected to UV–Visible irradiation in a photostability test chamber for 10 days. The samples were analyzed for chemical changes by a specific stability-indicating high-performance liquid chromatography method. pH of the microenvironment was determined in 10% w/v aqueous slurry of the salts. The observed order of photostability was: prazosin hydrochloride anhydrous > prazosin camsylate anhydrous ∼ prazosin-free base > prazosin hydrochloride polyhydrate > prazosin tosylate anhydrous > prazosin oxalate dihydrate ∼ prazosin tosylate monohydrate. Multivariate analysis of the photodegradation behavior suggested predominant contribution of the state of hydration and also intrinsic photosensitivity of the counterion. Overall, hydrated salts showed higher photodegradation compared to their anhydrous counterparts. Within the anhydrous salts, aromatic and carbonyl counterion-containing salts showed higher susceptibility to light. The pH of microenvironment furthermore contributed to photodegradation of prazosin salts, especially for drug counterions with inherent higher pH. The study reveals importance of selection of a suitable drug salt form for photosensitive drugs during preformulation stage of drug development.  相似文献   

5.
Thermally induced transition between anhydrous and hydrated forms of highly crystalline beta-chitin was studied by differential thermal calorimetry (DSC) and X-ray diffraction. DSC of wet beta-chitin in a sealed pan gave two well-defined endothermic peaks at 85.2 and 104.7 degrees C on heating and one broad exothermic peak at between 60 and 0 degrees C on cooling. These peaks were highly reproducible and became more distinct after repeated heating-cooling cycles. The X-ray diffraction pattern of wet beta-chitin at elevated temperature showed corresponding changes in d-spacing between the sheets formed by stacking of chitin molecules. These phenomena clearly show that water is reversibly incorporated into the beta-chitin crystal and that the temperature change induces transitions between anhydrous, monohydrate, and dihydrate forms. The DSC behavior in heating-cooling cycles, including reversion between the two endothermic peaks, indicated that the transition between monohydrate and dihydrate was a fast and narrow-temperature process, whereas the one between the anhydrous and the monohydrate form was a slow and wide-temperature process.  相似文献   

6.
We have examined the DNA damage produced by reaction of peroxyl radicals with human fibroblast DNA. DNA damage consisted of both strand breaks and base modifications. The extent of strand breaks and base modifications induced as a function of peroxyl radical concentration was determined by quantitation of fragment size distributions using denaturing glyoxal-agarose gel electrophoresis. Both strand breaks and base modifications increased in a log linear fashion with respect to peroxyl radical concentration. Oxidative base modifications were observed to occur to a greater extent than strand breaks at every concentration measured. The sequence-specific distribution of peroxyl radical induced base damage was mapped for 803 nucleotide positions using the method of ligation mediated PCR. A total of 87% of all guanine positions in the examined sequences was found to be significantly oxidized. The order of reactivity of DNA bases toward oxidation by peroxyl radicals was found to be G > C > T. Adenine is essentially unreactive. The yield of oxidative base modifications at guanines and cytosines by peroxyl radicals depends on the exact specification of 5' and 3' flanking bases in a polarity dependent manner. Every guanine in the 5'XGC3' motif was found to be oxidized, where X is any 5' neighbor. In contrast, 5' and 3' purine flanks drastically reduced the extent of peroxyl radical G oxidation. The pattern of base modification and the influence of nearest neighbors differs substantially from that previously reported for hydrogen peroxide damage mediated by low valent transition metal ions for the identical DNA sequences.  相似文献   

7.
Nitroxides are cell-permeable stable radicals that protect biomolecules from oxidative damage in several ways. The mechanisms of protection studied to date include removal of superoxide radicals as SOD-mimics, oxidation of transition metal ions to preempt the Fenton reaction, and scavenging carbon-centered radicals. However, there is no agreement regarding the reaction of piperidine nitroxides with peroxyl radicals. The question of whether they can protect by scavenging peroxyl radicals is important because these radicals are formed in the presence of oxygen abundant in biological tissues. To further our understanding of the antioxidative behavior of piperidine nitroxides, we studied their effect on biochemical systems exposed to the water soluble radical initiator 2,2'-azobis (2-amidinopropane) hydrochloride (AAPH). AAPH thermally decomposes to yield tert-amidinopropane radicals (t-AP(*)) that readily react with oxygen to form peroxyl radicals (t-APOO(*)). It has recently been reported that piperidine nitroxides protect plasmid DNA from t-AP(*) though not from t-APOO(*). The present study was directed at the question of whether these nitroxides can protect biological systems from damage inflicted by peroxyl radicals. The reaction of nitroxides with AAPH-derived radicals was followed by cyclic voltammetry and electron paramagnetic resonance spectroscopy, whereas the accumulation of peroxide was iodometrically assayed. Assaying DNA damage in vitro, we demonstrate that piperidine nitroxides protect from both t-AP(*) and t-APOO(*). Similarly, nitroxides inhibit AAPH-induced enzyme inactivation. The results indicate that piperidine nitroxides protect the target molecule by reacting with and detoxifying peroxyl radicals.  相似文献   

8.
Single crystals of anhydrous thymine (Ta) and cytosine monohydrate (Cm) were irradiated at room temperature using X rays, 20 MeV protons and 35 MeV alpha particles, and the relative distributions of the various radiation-induced stable radicals in the crystals were investigated. These two crystal systems were chosen because of systematic differences in their molecular packing and hydrogen-bonding network. The radicals stabilized in these systems have previously been identified and analyzed by several authors. Experimental EPR spectra could thus be reconstructed based on simulated benchmark spectra using a fitting procedure, yielding the relative amounts of the different radical species. It was found that the relative amounts of a given radical species varied with the type of radiation used, with differences being most prominent between alpha particles and protons or X rays. In Ta, an increased production of hydrogen addition or abstraction radicals was found after exposure to alpha particles. These radicals are believed to be formed predominantly from superexcited states, resulting from the higher density of ionizations along the track of the alpha particle. A corresponding reduction in radicals derived from proton transfer reactions was observed in comparison with exposure to protons or X rays. In Cm, the differences were smaller than for Ta. This is probably due to differences in the arrangement of the crystal lattice between the two systems, with Cm having an extended hydrogen-bonding network promoting fast proton transfers after ionization. Most interesting, however, was the observation that more than 50% of the radical population in Cm not could be accounted for by known radical species. The population with an unknown origin was represented with a broad singlet, having a line width of 1.5 mT and a g value of 2.0045, in the spectral reconstruction procedure.  相似文献   

9.
Free radicals from the one-electron reduction of adrenochrome have been studied in aqueous solutions. These radicals have been detected and identified by electron spin resonance spectroscopy, using spin stabilization methods (complexation with diamagnetic metal ions) to enhance radical concentrations. It is shown that the radicals have a characteristic ESR spectrum enabling their identification in complex systems. The spin density distribution in the radicals has been studied as a function of complexing metal ions and solvent composition. In the presence of oxidants (e.g., oxygen) the spectrum of the radical is replaced by that derived from the one-electron exidation of adrenochrome.  相似文献   

10.
It has been suggested that carbonate radical anions are biologically important because they may be produced during the inflammatory response. The carbonate radicals can selectively oxidize guanine in DNA and RNA by one-electron transfer mechanisms and the guanine radicals thus formed decay by diverse competing pathways with other free radicals or nucleophiles. Using a photochemical method to generate CO(3)(-) radicals in vitro, we compare the distributions of products initiated by the one-electron oxidation of guanine in the trinucleotides 5'-r(GpCpU) and 5'-d(GpCpU) in aqueous buffer solutions (pH 7.5). Similar distributions of stable end products identified by LC-MS/MS methods were found in both cases. The guanine oxidation products include the diastereomeric pair of spiroiminodihydantoin (Sp) and 2,5-diamino-4H-imidazolone (Iz). In addition, intrastrand cross-linked products involving covalent bonds between the G and the U bases (GCU) were also found, although with different relative yields in the 2'-deoxy- and the ribotrinucleotides. The positive-ion MS/MS spectra of the 5'-r(GpCpU) and 5'-d(GpCpU) products clearly indicate the presence of covalently linked G-U products that have a mass smaller by 2 Da than the sum of the G and U bases in both types of trinucleotides. The 5'-d(GCU) cross-linked product was further characterized by 1D and 2D NMR methods that confirm its cyclic structure in which the guanine C8 atom is covalently linked to the uracil N3 atom.  相似文献   

11.
In living tissues under inflammatory conditions, superoxide radicals (O(2)*)) are generated and are known to cause oxidative DNA damage. However, the mechanisms of action are poorly understood. It is shown here that the combination of O(2)* with guanine neutral radicals, G(-H)* in single- or double-stranded oligodeoxyribonucleotides (rate constant of 4.7 +/- 1.0 x 10(8) m(-1) s(-1) in both cases), culminates in the formation of oxidatively modified guanine bases (major product, imidazolone; minor product, 8-oxo-7,8-dihydroguanine). The G(-H)* and O(2)* radicals were generated by intense 308 nm excimer laser pulses resulting in the one-electron oxidation and deprotonation of guanine in the 5'-d(CC[2AP]-TCGCTACC) strands and the trapping of the ejected electrons by molecular oxygen (Shafirovich, V., Dourandin, A., Huang, W., Luneva, N. P., and Geacintov, N. E. (2000) Phys. Chem. Chem. Phys. 2, 4399-4408). The addition of Cu,Zn-superoxide dismutase, known to react rapidly with superoxide, dramatically enhances the life-times of guanine radicals from 4 to 7 ms to 0.2-0.6 s in the presence of 5 microm superoxide dismutase. Oxygen-18 isotope labeling experiments reveal two pathways of 8-oxo-7,8-dihydroguanine formation including either addition of O(2)* to the C-8 position of G(-H)* (in the presence of oxygen), or the hydration of G(-H)* (in the absence of oxygen). The formation of the guanine lesions via combination of guanine and superoxide radicals is greatly reduced in the presence of typical antioxidants such as trolox and catechol that rapidly regenerate guanine by the reductive "repair" of G(-H)* radicals. The mechanistic aspects of the radical reactions that either regenerate undamaged guanine in DNA or lead to oxidatively modified guanine bases are discussed.  相似文献   

12.
The possibility to amorphize anhydrous crystalline sugars, like lactose, trehalose and glucose, by mechanical milling was previously reported. We test here the possibility to amorphize the corresponding crystalline hydrates: lactose monohydrate, trehalose dihydrate and glucose monohydrate using fully identical milling procedures. The results show that only the first hydrate amorphizes while the other two remain structurally invariant. These different behaviours are attributed to the plasticizing effect of the structural water molecules which can decrease the glass transition temperature below the milling temperature. The results reveal clearly the fundamental role of the glass transition in the solid-state amorphization process induced by milling, and they also explain why crystalline hydrates are systematically more difficult to amorphize by milling than their anhydrous counterpart. The investigations have been performed by differential scanning calorimetry and powder X-ray diffraction.  相似文献   

13.
Hydrogen peroxide-induced base damage in deoxyribonucleic acid   总被引:5,自引:0,他引:5  
Aqueous solutions of calf thymus deoxyribonucleic acid (DNA) were exposed to hydrogen peroxide in the presence of air. Base products formed in DNA were identified and quantitated following acid hydrolysis and trimethylsilylation using gas chromatography-mass spectrometry. The yields of these products were dependent upon the hydrogen peroxide concentration, and increased in the following order: 8-hydroxyadenine, cytosine glycol, 2,6-diamino-4-hydroxy-5-formamidopyrimidine, 8-hydroxyguanine, thymine glycol, and 4,6-diamino-5-formamidopyrimidine. Previous studies have shown that these compounds are typically formed in DNA in aqueous solution by hydroxyl radicals generated by ionizing radiation. Hydrogen peroxide is thought to participate in a Fenton-like reaction with transition metals, which are readily bound to DNA in trace quantities, resulting in the production of hydroxyl radicals close to the DNA. This proposed mechanism was examined by exposing DNA to hydrogen peroxide either in the presence of a hydroxyl radical scavenger or following pretreatment of DNA with metal-ion chelators. The results indicate that trace quantities of transition metal ions can react readily with hydrogen peroxide to produce radical species. The production of radical species was monitored by determining the altered bases that resulted from the reaction between radicals and DNA. The yields of the base products were reduced by 40 to 60% with 10 mmol dm-3 of dimethyl sulfoxide. A 100-fold increase in the concentration of dimethyl sulfoxide did not result in a further reduction in hydrogen peroxide-induced base damage. DNA which was freed from bound metal ions by pretreatment with metal ion chelators followed by exhaustive dialysis was found to be an ineffective substrate for hydrogen peroxide. The yields of base products measured in this DNA were at background levels. These results support the role of metal ions bound to DNA in the site-specific formation of highly reactive radical species, most likely hydroxyl radicals, in hydrogen peroxide-induced damage to the bases in DNA.  相似文献   

14.
Radical formation and hole transfer were investigated in crystals of cytosine.HCl (C.HCl) doped with 0-1.1 mol-% 5-methylcytosine x HCl (5MC x HCl). The doping level was determined by NMR spectroscopy. Crystals and polycrystalline samples were X-irradiated at 295 K, 77 K and 12 K and studied with EPR, ENDOR and FSE spectroscopy at these temperatures. At 295 K the dominant radicals were the so-called 3alphaH radical, formed in 5MC by a net H-abstraction from the methyl group, and the cytosine C6 H-addition (5-yl) radical. At 12 K five radicals were identified. These were the 3alphaH radical, cytosine reduction and oxidation products, and the cytosine C6 and C5 H-addition (5-yl and 6-yl, respectively) radicals. The spectroscopic parameters for the 3alphaH radical are very similar to those of a radical observed previously in the crystalline cytosine derivatives cytidine (CR), 2'deoxycytidine hydrochloride (CdR x HCl), 5'dCMP and 3'CMP as well as in the uracil derivative 2-thiouracil (2-TU). It was shown that amounts of the order of tenths of a percent 5MC x HCl doped into crystals of C.HCl give rise to a considerable yield of 3alphaH radicals after exposure to ionizing radiation both at room temperature and at lower temperatures. This supports a previous suggestion that naturally occurring 5-methylated cytosine impurities may be responsible for the formation of 3alphaH radicals in the crystalline cytosine derivatives CR, CdR.HCl, 5'dCMP and 3'CMP and suggests that the 3alphaH radical in these systems is a 5-methylated base-centered radical. The total radical yield in doped C x HCl crystals increased considerably with the doping level, both at low temperatures and at room temperature, implying that the 3alphaH radical is more stable than the primary cytosine radicals. The relative amounts of the 3alphaH radical were obtained by using simulated benchmark spectra to reconstruct experimental EPR spectra of doped polycrystalline samples. Evidence is presented suggesting that the enhanced yield of the 3alphaH radical in doped samples is due to holes originally formed at cytosine bases and transferred to 5-methylcytosine bases in addition to the 3alphaH radical being less exposed to recombination than other cytosine radicals.  相似文献   

15.
Free radicals do not commonly add to nucleotides in DNA, despite the fact that radicals are produced in all aerobically metabolizing cells. Why is this? For oxy-radicals, the ratio of the rate constant for addition to double bonds divided by that for H-abstraction from good H-donors parallels the electrophilicity of the radical, and among oxy-radicals the hydroxyl radical is the most electrophilic, with an unusually high ratio of Kad/kH. The hydroxyl radical also is very reactive in H-atom abstraction reactions, with a large absolute value of kH. However, the hydroxyl radical's high reactivity makes it unselective and relatively nondiscriminating between H-abstraction from a sugar moiety in DNA and penetration to, and reaction with, a base. Oxy-radicals such as alkoxyl and peroxyl radicals do not have as high electrophilicity or as high reactivity. Interestingly, carbon-centered radicals (such as the methyl radical) also can both add to double bonds and abstract H-atoms, but carbon-centered radicals are not commonly observed to add to DNA bases. However, they cannot be generated near DNA in vivo. In contrast, hydroxyl radical generating systems appear to complex with DNA and produce the hydroxyl radical in the immediate vicinity of the DNA, producing a type of DNA damage that is called site specific. Thus, addition of a radical to a DNA base may require all three features possessed by the hydroxyl radical: high electrophilicity, high thermokinetic reactivity, and a mechanism for production near DNA.  相似文献   

16.
Radical formation in single crystals of hypoxanthine.HCl.H2O, inosine and Na2-5'-IMP.(7.5 H2O) by X-irradiation has been studied using electron-spin-resonance spectroscopy at 9.5 and 35 GHz. In all crystals both H-addition radicals at position C2 and C8 of the purine ring are found. The coupling constants of these two radicals are different and depend strongly on the protonation state of the base. INDO-calculations indicate that the C8-radical is protonated at O6. In Na2-5'-IMP OH-addition radicals at position C2 of the purine ring are formed. Electron adduct radicals are found in the neutral and the N7-protonated base after X-irradiation at 77 K. In Na2-5'-IMP no electron adduct is formed but a radical which probably is the cation. In hypoxanthine.HCl.H2O a radical could be observed after X-irradiation at 77 K, which results from addition of a Cl- to the nitrogen N1.  相似文献   

17.
DNA sequence context has emerged as a critical determinant of the location and quantity of nucleobase damage caused by many oxidizing agents. However, the complexity of nucleobase and 2-deoxyribose damage caused by strong oxidants such as ionizing radiation and the Fenton chemistry of Fe2+-EDTA/H2O2 poses a challenge to defining the location of nucleobase damage and the effects of sequence context on damage chemistry in DNA. To address this problem, we developed a gel-based method that allows quantification of nucleobase damage in oxidized DNA by exploiting Escherichia coli exonuclease III to remove fragments containing direct strand breaks and abasic sites. The rigor of the method was verified in studies of guanine oxidation by photooxidized riboflavin and nitrosoperoxycarbonate, for which different effects of sequence context have been demonstrated by other approaches (Margolin, Y., Cloutier, J. F., Shafirovich, V., Geacintov, N. E., and Dedon, P. C. (2006) Nat. Chem. Biol. 2, 365-366). Using duplex oligodeoxynucleotides containing all possible three-nucleotide sequence contexts for guanine, the method was used to assess the role of DNA sequence context in hydroxyl radical-induced guanine oxidation associated with gamma-radiation and Fe2+-EDTA/H2O2. The results revealed both differences and similarities for G oxidation by hydroxyl radicals and by one-electron oxidation by riboflavin-mediated photooxidation, which is consistent with the predominance of oxidation pathways for hydroxyl radicals other than one-electron oxidation to form guanine radical cations. Although the relative quantities of G oxidation produced by hydroxyl radicals were more weakly correlated with sequence-specific ionization potential than G oxidation produced by riboflavin, damage produced by both hydroxyl radical generators and riboflavin within two- and three-base runs of G showed biases in location that are consistent with a role for electron transfer in defining the location of the damage products. Furthermore, both gamma-radiation and Fe2+-EDTA/H2O2 showed relatively modest effects of sequence context on the proportions of different damage products sensitive to E. coli formamidopyrimidine DNA glycosylase and hot piperidine, although GT-containing sequence contexts displayed subtle biases in damage chemistry (formamidopyrimidine DNA glycosylase/piperidine ratio). Overall, the results are consistent with the known chemistry of guanine oxidation by hydroxyl radical and demonstrate that charge migration plays a relatively minor role in determining the location and chemistry of hydroxyl radical-mediated oxidative damage to guanine in DNA.  相似文献   

18.
Guanine oxidation by electron transfer results in the formation of a guanine radical cation, which is at the origin of long-range charge transport through double-stranded DNA. It is possible to observe guanine lesions at a long distance from the oxidative reagent covalently bound to DNA owing to the migration of the positive hole in the DNA pi-stacks. This phenomenon of long-range hole transport is classically studied in the literature with photosensitizers used as one-electron oxidants. It is shown in the present work that the process of long-range charge transport and the concomitant formation of guanine lesions at a long distance can be observed also in the case of two-electron oxidants. This is the signature of the formation of a transient guanine radical cation in the course of the two-electron abstraction process and consequently evidence of the separated one plus one electron abstraction steps. Long-range charge transport is likely to be a universal mechanism for any two-electron oxidant acting by electron abstraction provided that the second electron abstraction is slower than hole transfer.  相似文献   

19.
In a study of guanine.HCl.H2O (Gm) single crystals X-irradiated at temperatures between 10 and 77 K, three radical species were found and characterized by ESR and ENDOR spectroscopy. All three are primary products in that they were present immediately following irradiation at T less than 10 K. Radical I, which apparently can exist in two slightly different conformations, was identified as the product of electron gain by the parent molecule and subsequent protonation at O6. Radical I decayed only after warming the crystals beyond 250 K. Radical II was the guanine cation previously reported (D. M. Close, E. Sagstuen, and W. H. Nelson, J. Chem. Phys. 82, 4386 (1985)); however, ENDOR data are reported here which confirm the previous results. The guanine cation in Gm resulted from electron loss from the parent and subsequent deprotonation at N7. It is proposed that Radical III results from OH attack at C8 of the parent molecule, followed by rupture of the C8-N9 bond and ring opening. The OH radicals thought to produce Radical III result from electron loss by the cocrystallized water molecules. The reaction leading to Radical III, unusual in solid-state radiation chemistry, is thought to be mediated by the specific hydrogen bonding network in this crystal.  相似文献   

20.
Solid pyrimidine nucleic acid bases (cytosine, thymine, and uracil) were gamma-irradiated (50 KGy) and dissolved in deaerated solutions of adriamycin in water and dimethylsulfoxide (DMSO). Analogous experiments using unirradiated pyrimidines as controls were also performed. In water only gamma-irradiated cytosine showed a reaction with the adriamycin yielding a single ESR peak (g = 2.0033) consistent with the adriamycin semiquinone radical. Since the unirradiated cytosine gave no reaction, the result suggests an electron transfer from cytosine radicals (generated by gamma-radiolysis) to adriamycin. In DMSO the three gamma-irradiated and unirradiated pyrimidines reacted with adriamycin yielding the adriamycin semiquinone radical observed by ESR. These results suggest that in DMSO an electron is transferred to adriamycin from the pyrimidine radicals and from the parent pyrimidine molecules. However, the process is on the order of 10(5) times more efficient for the pyrimidine radicals. Superoxide radicals (O2-.) were formed following addition of oxygen to the deaerated DMSO solutions containing adriamycin semiquinone radicals. O2-. was spin trapped using 5,5-dimethyl-1-pyrroline-N-oxide (DMPO). The results show a possible reaction sequence in which an electron transferred to adriamycin, by pyrimidine radicals and parent pyrimidine molecules, is subsequently transferred to dissolved oxygen.  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