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.
D J Kaplan  L H Hurley 《Biochemistry》1981,20(26):7572-7580
Anthramycin and mitomycin C (MC) are two DNA reactive drugs, which bind covalently to GC pairs producing different effects on DNA: anthramycin stiffening and MC distorsion. This paper describes experiments in which we have used anthramycin as a probe to sense quantitatively the effects on DNA of MC binding. Saturation binding experiments show that both anthramycin and MC partially inhibit the binding of the other drug to DNA (maximum inhibition by MC and anthramycin, 22.4% and 19.7%, respectively) but by a mechanism other than direct site exclusion. This suggests that MC binds in the major groove of DNA, since anthramycin is known to bind in the minor groove. An abrupt reduction in the binding of anthramycin to DNA-MC complexes occurs between MC binding ratios of 0.030 and 0.035, which parallels and probably results from sudden intensification of a MC-induced DNA conformational change occurring between these binding ratios. Dialysis measurements indicate that anthramycin is very possibly binding at sites distant from MC sites and suggest a clustering of closely bound MC chromophores resulting from possible cooperative binding. S1 nuclease digest experiments demonstrate an initial enhancement of nuclease activity in DNA-MC complexes, the magnitude of which correlates well with the reduction of anthramycin binding, relative to the degree of MC binding. The enhanced nuclease activity in these complexes indicates regions of exposed DNA or helix base distortion which is related to or is the result of conformational change.  相似文献   

2.
The binding constants for interaction of the anticancer agents mitoxantrone and ametantrone and several congeners with calf thymus DNA and the effects of ionic strength changes have been determined spectrophotometrically. The agents show a preference for certain sequences, particularly those with GC base pairs, and the magnitude of the specificity depends on the specific substituents on the anthraquinone ring system. The binding constant for mitoxantrone with calf thymus DNA in 0.1 M Na+, pH 7, is approximately 6 X 10(6) M-1, and the rate constant for the sodium dodecyl sulfate driven dissociation of mitoxantrone from its calf thymus DNA complex under the same solution conditions and 20 degrees C was determined to be 1.3 s-1. The unwinding angle of mitoxantrone determined independently by viscosity measurements and by a novel assay employing calf thymus topoisomerase shows excellent agreement for a value of 17.5 degrees. The viscosity increase of sonicated calf thymus DNA varies considerably with the substituent on the anthraquinone ring system. Binding studies employing T4 and phi w-14 DNAs in which the major groove is occluded and the reverse experiment with anthramycin-treated calf thymus DNA indicate at least part of the mitoxantrone molecule may lie in the minor groove.  相似文献   

3.
The intrinsic viscosity of sonicated calf thymus DNA (molecular weight 4–5 × 105) increases and the sedimentation constant decreases, with increasing binding of proflavine at 0. 2 ionic strength and at 25°C. The measurements correspond to a linear increase in length of the almost rodlike DNA molecules with the amount of proflavine bound; independent calculations from viscosity and sedimentation measurements yield almost identical results. Over the range of r (moles of proflavine bound per moles of nucleotides) equal to zero to r = 0.13, the length increases by about 20%. This extension is compatible with the intercalation hypothesis proposed by Lerman. Density increments at various values of r, at constant chemical potential of diffusible solutes, were determined. It was also found that, in addition to the known isosbestic point of DNA-proflavine complexes at 455.5 mμ, an additional isosbestic point exists at 225.5 mμ; this proved extremely useful for the evaluation of binding studies.  相似文献   

4.
Two novel antibiotics were isolated, designated compounds 1QN and 2QN respectively, having quinoline rings in place of one or both of the quinoxaline chromophores of echinomycin. Each removes and reverses the supercoiling of closed circular duplex DNA from bacteriophage PM2 in the fashion characteristic of intercalating drugs. For compound 1QN, the unwinding angle at I0.01 is almost twice that of ethidium, whereas for compound 2QN the value is indistinguishable from that of ethidium. Binding of both analogues produced changes in the viscosity of sonicated rod-like DNA fragments corresponding to double the helix extension found with ethidium, a feature characteristic of bifunctional intercalation by quinoxaline antibiotics. These results suggest that both compounds 1QN and 2QN behave as bifunctional intercalators but that compound 2QN produces only half the helix unwinding seen with compound 1QN and the natural quinoxalines. Binding curves for the interaction of both analogues with a variety of synthetic and naturally occurring nucleic acids were determined by solvent-partition analysis. Values for compound 2QN were also obtained by a fluorimetric method and found to agree well with the solvent-partition measurements. Compound 1QN bound most tightly to Micrococcus lysodeikticus DNA and, like echinomycin, exhibited a broad preference for (G + C)-rich DNA species. For compound 2QN no marked (G + C) preference was indicated, and the tightest binding among the natural DNA species studied was found with DNA from Escherichia coli. The two analogues also displayed different patterns of specificity in their interaction with synthetic nucleic acids. Compound 2QN bound to poly(dA-dT) slightly more tightly than to poly-(dG-dC), whereas compound 1QN displayed a large (approx. 11-fold) preference in the opposite sense. There was evidence of co-operativity in the binding to poly(dA-dT). It may be concluded that the chromophore moieties play an active role in determining the capacity of quinomycin antibiotics to recognize and bind selectively to specific sequences in DNA.  相似文献   

5.
The intrinsic, viscosities at zero shear rate of defined complexes of proflavine, 9-aminoacridine, and 9-amino-l,2,3,4-tetrahydroaeridine with calf thymus DNA have been determined at, various ionic strengths by means of rotating cylinder viscometers. By controlled adjustment, of the composition of the mixtures, the amount of bound acridine (r moles/g.-atom DNA phosphorus) was maintained constant at different dilutions. The intrinsic viscosities of the complexes increased with r up to r values (ca. 0.16–0.20) corresponding to the end of the process of strong binding of the acridinium cations. However, complex formation between the acridines and thermally denatured DNA caused either a marked decrease in viscosity (at the low ionic strengths of 0.0015 and 0.005) or no change at all (ionic strength 0.1). These results are discussed in the light of presently available hydrodynamic theories relating the intrinsic, viscosity of DNA to its molecular extension.  相似文献   

6.
The binding of echinomycin to deoxyribonucleic acid.   总被引:20,自引:4,他引:16       下载免费PDF全文
Echinomycin is a peptide antibiotic which binds strongly to double-helical DNA up to a limit of approximately one molecule per five base-pairs. There is no detectable interaction with rRNA and only extremely feeble non-specific interaction with poly(rA)-poly(rU). Heat denaturation of DNA greatly decreases the binding, and similarly limited interaction is observed with naturally occurring single-stranded DNA. Association constants for binding to nine double-helical DNA species from different sources are presented; they vary by a factor of approximately 10, but are not simply related to the gross base composition. The interaction with DNA is ionic-strength-dependent, the binding constant falling by a factor of 4 when the ionic strength is raised from 0.01 to 0.10mol/litre. From the effect of temperature on the association constant for calf thymus DNA, the enthalpy of interaction is calculated to be about -13kJ/mol (-3kcal/mol). Binding of echinomycin persists in CsCl gradients and the buoyant density of nicked bacteriophage PM2 DNA is decreased by 25 mg/ml. Echinomycin interacts strongly with certain synthetic poly-deoxynucleotides, the binding constant decreasing in the order poly(dG)-poly(dC) greater than poly(dG-dC) greater than poly(dA-dT). For the latter two polymers the number of base-pairs occluded per bound antibiotic molecule is calculated to be three, whereas for poly(dG)-poly(dC) it is estimated to be four to five. Poly(dA)-poly(dT) and poly(dI)-poly(dC) interact only very weakly with the antibiotic. Poly(dI-dC) interacts to a slightly greater extent, but the binding curve is quite unlike that seen with the three strongly binding synthetic polynucleotides. Echinomycin affects the supercoiling of closed circular duplex bacteriophage PM2 DNA in the characteristic fashion of intercalating drugs. At low ionic strength the unwinding angle is almost twice that of ethidium. Likewise the extension of the helix, determined from changes in the viscosity of rod-like sonicated DNA fragments, is nearly double that expected for a simple (monofunctional) intercalation process. On this basis the interaction process is characterized as bifunctional intercalation. At higher ionic strength the unwinding angle relative to that of ethidium and the helix extension per bound echinomycin molecule fall, indicating a smooth progression towards more nearly monofunctional intercalation. Two simpler compounds which act as analogues of the quinoxaline chromophores of echinomycin, quinoxaline-2-carboxamide and the trypanocidal drug Bayer 7602, interact with DNA very much more weakly than does echinomycin, showing that the peptide portion of the antibiotic plays an essential role in determining the strength and specificity of the interaction.  相似文献   

7.
Quinomycin C, triostin A and triostin C are peptide antibiotics of the quinoxaline family, of which echinomycin (quinomycin A) is also a member. They all remove and reverse the supercoiling of closed circular duplex DNA from bacteriophage PM2 in the fashion characteristic of intercalating drugs, and the unwinding angle at I 0.01 is, in all cases, almost twice that of ethidium. Thus, as with echinomycin, they can be characterized as bifunctional intercalating agents. For the triostins this conclusion has been confirmed by measurements of changes in the viscosity of sonicated rod-like DNA fragments; the helix extension was found to be almost double that expected for a simple monofunctional intercalation process. For triostin A, further evidence for bifunctionality was derived from the cross-over point of binding isotherms to nicked circular and closed circular bacteriophage-PM2DNA. Binding curves for the interaction of quinomycin C and triostin A with a variety of synthetic and naturally occurring nucleic acids were determined by solvent-partition analysis, but triostin C was too insoluble in aqueous solution to make this method applicable. For quinomycin C the highest binding constant was found with Micrococcus lysodeikticus DNA, and its pattern of specificity among natural DNA species was broadly similar to that of echinomycin, although the binding constants were 2--6 times as large. For triostin A the highest binding constant was again found for M. lysodeikticus DNA, but the specificity pattern was quite different from that of the quinomycins. In particular, triostin A bound better to poly(dA-dT) than to the poly(dG-dC) whereas this order was reversed for quinomycin C. There was also evidence that the binding to poly(dA-dT) might be co-operative in nature. No significant interaction could be detected with poly(dA).poly(dT) or with RNA from Escherichia coli. Poly(dG).poly(dC) gave variable results, depending on the source of the polymer. The different patterns of specificity displayed by the quinomycins and triostins are tentatively ascribed to differences in their conformations in solution.  相似文献   

8.
Lysates containing folded chromosomes of competent Bacillus subtilis were prepared. The chromosomes were supercoiled, as indicated by the biphasic response of their sedimentation rates to increasing concentrations of ethidium bromide. Limited incubation of the lysates with increasing concentrations of ribonucleases resulted in a gradual decrease in the sedimentation velocity of the deoxyribonucleic acid (DNA) until finally a constant S value was reached. Incubation with sonicated, 4,5',8-trimethylpsoralen-monoadducted, denatured, homologous donor DNA molecules at 37 degrees C and concomitant irradiation with long-wave ultraviolet light of the nucleoid-containing lysates resulted in the formation of complexes of the donor DNA molecules and the recipient chromosomes. This complex formation was stimulated when nucleoids were previously (i) unfolded by ribonuclease incubation, (ii) (partially) relaxed by X irradiation, or (iii) subjected to both treatments. Monoadducts were not essential. On the other hand, the complex-forming capacity of recipient chromosomes previously cross-linked by 4,5',8-trimethylpsoralen diadducts was greatly reduced, suggesting that strand separation of the recipient molecule was involved in the formation of the complex. None of these effects has been observed when heterologous (Escherichia coli) donor DNA has been used. When the same kind of experiments were carried out at 70 degrees C, donor-recipient DNA complexes were also formed and required strand separation and homology similar to donor-recipient complex formation at 37 degrees C. However, in contrast to what was found at 37 degrees C, unfolding plus relaxation of the nucleoids, as well as the absence of monoadducts in the donor DNA fragments, resulted in a decrease in complex formation. On the basis of these results, we assume that superhelicity can promote the in vitro assimilation of single-stranded donor DNA fragments by nucleoids of competents B. subtilis cells at 70 degrees C, but that at 37 degrees C a different mechanism is involved.  相似文献   

9.
The ultrasonic velocity at 3 MHz and the density in the nonsonicated and sonicated liposomes of dipalmitoylphosphatidylcholine have been measured in the temperature range from 0 degrees C to 55 degrees C. The results indicate that nonsonicated multilamellar vesicles undergo a weak first order transition which is analogous to the nematic-isotropic transition of liquid crystals. A sharp change in the ultrasonic velocity associated with the first order transition disappears when the multilamellar vesicles are sonicated. The bulk modulus of the lipid bilayer calculated from the ultrasonic velocity and the density of sonicated liposomes has a value of 3.0 X 10(10) dyne/cm2 at 20 degrees C, reaches a minimum value of 2.1 X 10(10) dyne/cm2 at its transition temperature and increases slightly to 2.2 X 10(10) dyne/cm2 at 50 degrees C.  相似文献   

10.
D G Gorenstein  K Lai 《Biochemistry》1989,28(7):2804-2812
31P NMR provides a convenient monitor of the phosphate ester backbone conformational changes upon binding of the intercalating drugs ethidium, quinacrine, and daunomycin to sonicated poly(A).poly(U) and calf thymus DNA. 31P chemical shifts can also be used to assess differences in the duplex unwinding angles in the presence of the drug. Thus a new 31P signal, 1.8-2.2 ppm downfield from the double-stranded helix signals, is observed in the ethidium ion-poly(A).poly(U) complex. This signal arises from phosphates which are in perturbed environments due to intercalation of the drug. This is in keeping with the hypothesis that the P-O ester torsional angle in phosphates linking the intercalated base pairs is more trans-like. Similar though smaller deshielding of the 31P signals is observed in sonicated poly(A).poly(U)-quinacrine complexes as well as in the daunomycin complexes. The effect of added ethidium ion, quinacrine, and daunomycin on the 31P spectra of sonicated calf thymus DNA is consistent with Wilson and Jones' (1982) earlier study. In these drug-DNA complexes the drug produces a gradual downfield shift in the DNA 31P signal without the appearance of a separate downfield peak. These differences are attributed to differences in the rate of chemical exchange of the drug between free and bound duplex states. The previous correlation of 31P chemical shift with drug duplex unwinding angle (Wilson & Jones, 1982) is confirmed for both the RNA and DNA duplexes.  相似文献   

11.
A series of 100 experiments was completed to determine if DNA is capable of transforming the genotype of a murine lymphoma (P388) in cell culture. The test system was concerned with the transformation of cells from 8-azaguanine (AZG) sensitivity to resistance. By the use of this marker, it was determined that transformation by DNA did occur, and that the efficiency of transformation was greatly increased by sonication of the DNA. A statistical analysis of 100 experiments demonstrated that the increase in the number of resistant cells after treatment with sonicated resistant DNA (R-DNA) was statistically significant (χ2 > 4.25, 0.05 > p > 0.02) in 66% of the experiments. DNA from sensitive parental cells and DNA from other sources produced no effect while DNase and UV treatment abrogated effective transformation by either sonicated or nonsonicated R-DNA. RNase was without effect. Sucrose gradient analysis of sonicated and nonsonicated R-DNA demonstrated that the peaks which correspond to the highest specific transforming activity are not altered by sonication and do not coincide with the OD260 peaks, in spite of the fact that sonication shifted the peak of maximum OD260 to a slower sedimenting region of the gradient. The major portion, however, of the transforming material did shift after sonication to the slower sedimenting region of the gradient and did coincide with the OD260 peak. The hereditary stability of the transformed cells was established by cloning a representative number of transformants, growing them in the absence of AZG for an extended period and then testing their ability to grow in graded concentrations of AZG. In addition, DNA extracted from transformants successfully transformed sensitive cells.  相似文献   

12.
The mode of binding of 5,6-dihydroflavopereirine and sempervirine to DNA has been investigated by absorption spectrophotometry, circular and electric linear dichroism, fluorescence and fluorescence polarization, viscosity increase of sonicated linear DNA and circular DNA unwinding. Although the spectroscopic properties of both compounds bound to DNA resembled those reported in our previous study of DNA complexes with two other alkaloids, and observed with planar intercalating compounds, only sempervirine was able to unwind circular DNA. The latter drug however showed signs characteristic of aggregation at the surface of the polyion. The differences between the behaviours of the four alkaloids so far investigated by us are interpreted on the basis of different extent of penetration of the chromophore ring into the DNA helix.  相似文献   

13.
The binding of naphthyridinomycin (NAP) to deoxyribonucleic acid was investigated using radioisotope labeled antibiotic. Dithiothreitol (DTT) enhances complex formation in a concentration dependent fashion but was found to be slightly inhibitory at concentrations above 10 mM. [C3H3]-NAP-DNA complexes, formed in the presence or absence of reducing reagents, were stable to Sephadex G-25 chromatography and precipitation with ethanol, indicating a strong bond formed between the drug and DNA. Time course studies showed that the difference between the binding of activated and non-activated antibiotic was a DTT-dependent burst. This was followed by a second phase of binding which was similar in both the activated and non-activated antibiotics. The activation of the antibiotic by DTT was a reversible reaction at pH 7.9. The activated form at pH 5.0 was extremely stable and did not revert to the unactivated form even after an 8-h incubation period. Antibiotic-DNA complex formation was pH independent between pH 5.0 and 7.0 for activated NAP. The non-activated antibiotic bound to DNA much better at pH 5.0 than at physiological pH values. Release of antibiotic from complexes (as followed by long term dialysis) formed in the presence of DTT and at pH 5.0 was biphasic, suggesting that the drug can bind to DNA in more than one way. A constant rate of antibiotic release was observed at pH 7.9 with or without DTT. At pH 2.0 and pH 12.0, greater than 95% of the antibiotic is released from the complexes. Most of the acid released antibiotic is NAP while most of the base released antibiotic had decomposed to a more polar compound. NAP binds well to calf thymus DNA, poly(dG) . poly(dC), and T4 DNA but shows significantly less affinity for poly(dA) . poly(dT), poly(dA . dT) . poly(dA . dT), poly(dG), poly(dC), poly(dI) . poly(dC) or poly(dG . dC) . poly(dG . dC). This specificity of NAP for DNA is similar to that observed for the pyrrolo(1,4)benzodiazepine antibiotics and saframycin A and S; all of which bind to double stranded DNA through their carbinolamine or masked carbinolamine functionalities. Two mechanisms which can explain the need for activation of NAP are also proposed.  相似文献   

14.
We previously reported, using a coimmunoprecipitation assay, that the B form (PR-B) of the human progesterone receptor from T47D human breast cancer cells dimerizes in solution with the A receptor (PR-A) and that the extent of dimerization correlates with receptor binding activity for specific DNA sequences [DeMarzo, A.M., Beck, C.A., O?ate, S.A., & Edwards, D.P. (1991) Proc. Natl. Acad. Sci. U.S.A. 88, 72-76]. This suggested that solution dimerization is an intermediate step in the receptor activation process. The present study has tested the effects of the progesterone antagonist RU486 on solution dimerization of progesterone receptors (PR). As determined by the coimmunoprecipitation assay, RU486 binding did not impair dimerization of receptors; rather, the antagonist promoted more efficient solution dimerization than the progestin agonist R5020. This enhanced receptor dimerization correlated with a higher DNA binding activity for transformed receptors bound with RU486. RU486 has been shown previously to produce two other alterations in the human PR when compared with R5020. PR-RU486 complexes in solution exhibit a faster sedimentation rate (6 S) on salt-containing sucrose density gradients than PR-R5020 complexes (4 S), and PR-DNA complexes have a faster electrophoretic mobility on gel-shift assays in the presence of RU486. We presently show that the 6 S PR-RU486 complex is a receptor monomer, not a dimer. The increased sedimentation rate and increased mobility on gel-shift assays promoted by RU486 were also observed with recombinant PR-A and PR-B separately expressed in insect cells from baculovirus vectors. These results suggest that RU486 induces a distinct conformational change both in PR monomers in solution and in dimers bound to DNA. We also examined whether conformational changes in PR induced by RU486 would prevent a PR polypeptide bound to RU486 from heterodimerization with another PR polypeptide bound to R5020. To evaluate this, PR-A and PR-B that were separately bound to R5020 or RU486 in whole cells were mixed in vitro. PR-A-RU486 was capable of dimerization with PR-B-R5020, and this was demonstrated for heterodimers both formed in solution and bound to specific DNA. The capability to form heterodimers in vitro raises the possibility that the antagonist action of RU486 in vivo could in part be imposed in a dominant negative fashion through heterodimerization between one receptor subunit bound to an agonist and another bound to RU486.  相似文献   

15.
E Kuchinka  J Seelig 《Biochemistry》1989,28(10):4216-4221
The binding of melittin to nonsonicated bilayer membranes composed of 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine was studied with an ultracentrifugation assay and with 2H and 31P nuclear magnetic resonance. Melittin binding could best be described by a partition equilibrium with Kp = (2.1 +/- 0.2) X 10(3) M-1, measuring the binding isotherm in the concentration range of 0-100 microM melittin and taking into account electrostatic effects by means of the Gouy-Chapman theory. This partition coefficient is smaller than that deduced for small sonicated vesicles and attests to the tighter lipid packing in the nonsonicated bilayers. Deuterium magnetic resonance revealed a conformational change of the phosphocholine head group upon melittin binding. The quadrupole splittings of the alpha and beta segments of the choline head group varied linearly with the amount of bound melittin but in opposite directions; i.e., the alpha splitting decreased, and the beta splitting increased. This conformational change is not specific to melittin but is a response of the phosphocholine head group to positive membrane surface charges in general. Quantitatively, melittin is one of the most efficient head-group modulators, the efficiency per unit charge comparable to that of charged local anesthetics or hydrophobic ions.  相似文献   

16.
The interaction of mitoxantrone, ametantrone and their Pd(II) complexes with DNA have been studied using absorption and circular dichroism spectroscopy. We have shown that mitoxantrone forms with Pd(II) a complex in which two Pd(II) ions are bound to two molecules of drug (D1 and D2). One Pd(II) ion is bound to the two nitrogens of the side chain on C-5 of molecule D1 and to the two nitrogens of the side chain on C-5 of molecule D2, whereas the second Pd(II) ion is bound to the nitrogens of the side chain on C-8 of molecule D1 and of molecule D2. The same complex is formed between Pd(II) and ametantrone. The stability constants for these complexes are, respectively, beta M = (1.4 +/- 0.5).10(19) and beta A = (2.5 +/- 0.5).10(18). They display antitumor activity against P 388 leukemia which compares with that of the free drugs. Interactions of the free drugs with DNA have been studied. Mitoxantrone and ametantrone are not optically active by themselves. However, through interaction with DNA, there is an induction of optical activity within the electronic transitions of both drugs. At a nucleotide/drug molar ratio lower than about 5 a CD signal of the couplet type is observed, suggesting that there is a coupling between the pi----pi transitions of the molecules of drugs intercalated between the base pairs. This coupling disappears when the molar ratio is increased. The interactions of the Pd(II) complexes with DNA do not give rise to induction of optical activity within the electronic transition of the drugs, indicating that the presence of the metal ion prevents the intercalation of the drugs between the base pairs.  相似文献   

17.
The method of chromatography of cell lysates on the columns with hydroxyapatite (HAP) and the method of ultracentrifugation of cell lysates in neutral sucrose gradient were used to study the mutagen-induced repair activity of human cells HEp-2 noninfected and chronically infected with measles and rubella viruses in order to determine the sedimentation properties of complexes containing DNA. Gamma-radiation, bleomycin, 4-nitroquinoline-1-oxide, and mitomycin C were used as DNA damaging agents. It was shown that the chronic infectious process inhibited repair of DNA damages induced by 4-nitroquinoline-1-oxide and mitomycin C and did not influence repair of DNA lesions caused by gamma-radiation and bleomycin.  相似文献   

18.
Two cobalt (II) complexes containing a dipyrido[3,2-a:2',3'-c]phenazine (dppz) base with the general formulation [Co(dppz)(dmp)(2)]Cl(2), where dmp is 4,7-dimethyl-1,10-phenanthroline ligand (4,7-dmp) (1) and 2,9-dimethyl-1,10-phenanthroline ligand (2,9-dmp) (2) were synthesized and characterized. Binding interactions of these complexes with calf thymus DNA were investigated by emission, absorption, circular dichroism, and viscosity studies, and the effects of the positions of methyl substitutions in phenanthroline coligands were investigated. The DNA binding constants obtained from the absorption spectral titrations decrease in the order of 1?>?2, which is consistent with the trend in apparent emission enhancement of the complexes on binding to calf thymus DNA. These observations were supported by circular dichroism spectroscopy and viscosity measurements and reveal that DNA binding affinity of the complexes depends on the position of methyl groups on the phenanthroline ligands.  相似文献   

19.
Reduction of mitomycin C by NaBH4 or by NADPH in the presence of a cell extract followed by exposure to air results in the generation of H2O2. This phenomenon occurs not only with free mitomycin but also with mitomycin irreversibly bound to DNA. In view of these findings, the antibiotic activity of mitomycin was tested in two bacterial systems: a facultative aerobic bacterium grown in the presence or absence of oxygen and an obligate anaerobic bacterium. No oxygen effect could be demonstrated in either case in the growth-inhibitory and bactericidal activity of the drug. Nevertheless, the H2O2 generating capacity of mitomycin-DNA complexes inside the nucleus may play a role in the drug-induced biological damage to the genetic material of cells.  相似文献   

20.
Variations of the membrane filter technique for deoxyribonucleic acid (DNA) hybridizations were studied with respect to Streptomyces species. At the temperatures required for specific hybridization of DNA with the high melting temperature (Tm) characteristic of Streptomyces, large amounts (up to 97%) of filter-bound DNA became eluted, in all reaction mixtures studied, within 21 hr. In most solutions this leaching was increased by the presence of sheared denatured DNA. Incubation of DNA-loaded filters in a solution of 50% formamide containing 6× standard saline citrate, at 48 C for 40 hr, was judged to be the best set of conditions tested based on relatively good retention of immobilized DNA, very low hybridization with unrelated DNA of a similarly high Tm (from Sarcina lutea), and the formation of complexes similar in thermal stability to the native DNA. The expression of results as sheared DNA bound in relation to long-chain DNA retained is recommended when a high concentration of sheared DNA relative to immobilized DNA is used.  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