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.
2.
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   

3.
Twelve anthraquinones (AQ) were evaluated for their ability to potentiate the antiviral activity of poly r(A-U) using a human foreskin fibroblast-vesicular stomatitis virus bioassay in which the AQ was combined with 0.2 mM poly r(A-U) to produce an AQ/ribonucleotide ratio of 1/4. Poly r(A-U) and the AQ alone were not effective antiviral agents. Five of the twelve AQs tested, mitoxantrone, adriamycin, ametantrone, carminic acid and daunomycin, enhanced the antiviral activity of poly r(A-U) 9- to 13-fold. The interferon-inducing activity of the five active AQ/poly r(A-U) combinations was equal to the sum of the interferon-inducing activities of their constituents. These five AQs appear to potentiate the antiviral activity of poly r(A-U) without superinduction of interferon.  相似文献   

4.
The equilibrium binding of the antitumor compound DHAQ, or mitoxantrone [1,4-dihydroxy-5,8-bis[[2-[(2-hydroxyethyl)amino]ethyl]amino]-9,10- anthracenedione], to various DNAs has been examined by optical titration and equilibrium dialysis methods. At low r (bound drug/DNA base pair) values, r less than 0.03, DHAQ binds, in a highly cooperative manner, to calf thymus and Micrococcus lysodeikticus DNAs. The binding isotherms for the interaction of DHAQ with Clostridium perfringens DNA and poly(dA-dT).poly(dA-dT) exhibit a small positive slope at low r values, suggestive of cooperative binding. In contrast, the binding of DHAQ to poly(dG-dC).poly(dG-dC) shows no evidence of cooperative binding even at very low r values. At higher r values (r greater than 0.05), the binding of DHAQ to all the DNAs studied is characterized by a neighbor-exclusion process. A model is proposed to account for the two modes of binding exhibited in the cooperative binding isotherms. The main feature of the proposed model is that local sequence and structural heterogeneity of the DNA give rise to sets of binding sites to which DHAQ binds in a highly cooperative manner, while the majority of the DNA sites bind DHAQ via a neighbor-exclusion process. This two-site model reproduces the observed binding isotherms and leads to the conclusion that DHAQ binds in clusters to selected regions of DNA. It is suggested that clustering may play a role in the physiological activity of drugs.  相似文献   

5.
Dihydroxyanthraquinone (DHAQ) and ametantrone (anthraquinone) are two new anthracenedione antineoplastic agents which were found by proton NMR spectroscopy to self-associate in aqueous media. Self-association was consistent with a bimolecular model, with average association constant values of 3400 and 2900 m?1 determined for DHAQ and ametantrone, respectively. Both anthracenediones interacted with the flavin nucleotides FMN and FAD to produce concentration-dependent upfield shifts of the flavin isoalloxazine ring proton signals, as observed by proton NMR spectroscopy. Average association constant values obtained for FMN-DHAQ, FAD-DHAQ, FMN-metantrone, and FAD-ametantrone complexation were 5100, 2600, 4300, and 1600 m?1, respectively. Optical difference spectroscopy confirmed FMN-DHAQ complexation, which resulted in a hyperchromic, bathochromic shift of the DHAQ spectrum following addition of FMN. These results were consistent with the formation of a ππ bimolecular ring-stacking complex. Information obtained on anthracenedione self-association and complexation with flavins may be of consequence in the interpretation of anthracenedione-DNA binding data and flavoprotein-mediated anthracenedione metabolic activation.  相似文献   

6.
The effect of dihydroxyanthraquinone (DHAQ), a new antitumor drug, on mammalian chromosome replication was investigated using simian virus 40 (SV40) as a model system. The maximum effect of inhibition on viral DNA synthesis was observed within 30-40 min after the addition of the drug. The extent of inhibition of viral DNA synthesis appeared to be directly related to the number of viral replicons which interact with DHAQ molecules in vivo. No apparent strand breakage of SV40 DNA was observed in infected cells treated with DHAQ ranging from 0.3 to 10 microM. However, strand breakage was induced upon cell lysis presumably by released nuclease. Repair of the damaged SV40 chromosomes in vitro resulted in the synthesis of completed supercoiled SV40 DNA. This repair synthesis was mostly confined to the region containing the replication origin of SV40 DNA as judged by the digestion of DNA with restriction endonucleases HindII and HindIII. Since SV40 DNA sequences close to the origin of replication are not complexed with histones to form a nucleosome structure, the results suggested that DHAQ may disturb chromosome structure by interacting preferentially to the nucleosome-free regions and causing the aberrant gene duplication and expression.  相似文献   

7.
The sodium dodecyl sulfate driven dissociation reactions of daunorubicin (1), mitoxantrone (2), ametantrone (3), and a related anthraquinone without hydroxyl groups on the ring or side chain (4) from calf thymus DNA, poly[d(G-C)]2, and poly[d(A-T)]2 have been investigated by stopped-flow kinetic methods. All four compounds exhibit biphasic dissociation reactions from their DNA complexes. Daunorubicin and mitoxantrone have similar dissociation rate constants that are lower than those for ametantrone and 4. The effect of temperature and ionic strength on both rate constants for each compound is similar. An analysis of the effects of salt on the two rate constants for daunorubicin and mitoxantrone suggests that both of these compounds bind to DNA through a mechanism that involves formation of an initial outside complex followed by intercalation. The daunorubicin dissociation results from both poly[d(G-C)]2 and poly[d(A-T)]2 can be fitted with a single exponential function, and the rate constants are quite close. The ametantrone and 4 polymer dissociation results can also be fitted with single exponential curves, but with these compounds the dissociation rate constants for the poly[d(G-C)]2 complexes are approximately 10 times lower than for the poly[d(A-T)]2 complexes. Mitoxantrone also has a much slower dissociation rate from poly[d(G-C)]2 than from poly[d(A-T)]2, but its dissociation from both polymers exhibits biphasic kinetics. Possible reasons for the biphasic behavior with the polymers, which is unique to mitoxantrone, are selective binding and dissociation from the alternating polymer intercalation sites and/or dual binding modes of the intercalator with both side chains in the same groove or with one side chain in each groove.  相似文献   

8.
The anthracenedione antineoplastic agents mitoxantrone and ametantrone are potent inhibitors of basal and drug-stimulated lipid peroxidation in a variety of subcellular systems (Kharasch, E. D., and Novak, R. F. (1983) J. Pharmacol. Exp. Ther. 226, 500-506). The mechanism by which these compounds function as antioxidants has been investigated using enzymic and chemical systems. Mitoxantrone and ametantrone inhibited NADPH-cytochrome P-450 reductase- and xanthine oxidase-catalyzed conjugated diene formation from linoleic acid in a concentration-dependent manner with half-maximal inhibition achieved at approximately 0.5 microM anthracenedione. Inhibition of linoleic acid peroxidation was not attributable to a decrease in P-450 reductase activity, hydroxyl radical scavenging, or iron chelation by the anthracenediones. Nonenzymic fatty acid peroxidation was also inhibited by the anthracenediones. Linoleic acid oxidation initiated by superoxide (ferrous iron autoxidation) or by hydroxyl radicals (Fenton's reagent) was diminished by mitoxantrone and ametantrone after a brief delay, suggesting an effect subsequent to activated oxygen-dependent initiation. In contrast, linoleic acid oxidation initiated by iron-dependent hydroperoxide decomposition was inhibited immediately. Reinitiation of linoleic acid oxidation in an anthracenedione-inhibited system was accomplished only by superoxide generation, but not by fatty acid hydroperoxide decomposition. These results suggest the anthracenediones diminished neither oxygen radical formation nor oxygen radical-dependent initiation of peroxidation. Rather, inhibition of fatty acid peroxidation by mitoxantrone and ametantrone results from the inhibition of hydroperoxide-dependent initiation and propagation reactions.  相似文献   

9.
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   

10.
Results are reported of a pH-metric and spectroscopic (CD and ESR) study of the complexes formed between the pseudo-peptide 1-hydroxy-4-(Gly-His-Lys)-anthraquinone (Q-GHK) since, when complexed to copper ions, Q-GHK has been shown to be very effective in promoting the formation of free radicals and inducing DNA cleavage. Q-GHK forms very stable complexes with copper, the major species being bonded to three nitrogen donors in the coordination plane: an imidazole-N of the His residue and the peptide nitrogens of the Gly and His residues. This species is probably stabilized through bonding of the fourth planar coordination site of Cu(II) to the 9-anthraquinone oxygen. At high Q-GHK:copper ratios a second Q-GHK molecule is coordinated through its imidazole-N donor.  相似文献   

11.
Although DNA helicases play important roles in the processing of DNA, little is known about the effects of DNA-interacting ligands on these helicases. Therefore, the effects of a wide variety of DNA-binding ligands on the unwinding and ATPase reactions catalyzed by Escherichia coli DNA helicase II were examined. DNA minor groove binders and simple DNA intercalators did not inhibit helicase II. However, DNA intercalators, such as mitoxantrone and nogalamycin, which position functionalities in the major groove upon binding duplex DNA, were potent inhibitors of helicase II. To determine the mechanism by which mitoxantrone inhibited helicase II, the unwinding and DNA-dependent ATPase activities of helicase II were measured using a spectrum of double- and single-stranded DNA substrates. Using either a 71-base pair (bp) M13mp7 partially duplexed DNA substrate or a 245-bp bluntended, fully duplexed DNA substrate, the apparent Ki value for inhibition by mitoxantrone of both the unwinding and ATPase reactions was approximately 1 microM for both substrates, suggesting that the mechanism of inhibition of helicase II by mitoxantrone is the same for both substrates and requires the presence of double-stranded structure. To strengthen this conclusion, the ability of mitoxantrone to inhibit the DNA-dependent ATPase activity of helicase II was determined using two single-stranded substrates, poly(dT) and the 245-bp substrate after heat denaturation. Using either substrate, mitoxantrone inhibited the ATPase activity of helicase II far less effectively. Thus, these results indicate that the intercalation of mitoxantrone into double-stranded DNA, with accompanying placement of functionalities in the major groove, generates a complex that impedes helicase II, resulting in both inhibition of ATP hydrolysis and unwinding activity. Furthermore, we report here that DNA-binding ligands inhibit the unwinding activity of helicases I and IV and Rep protein from E. coli, demonstrating that the inhibition observed for helicase II is not unique to this enzyme.  相似文献   

12.
The effects of two anthracenedione derivatives on in vitro cell survival and DNA of Chinese hamster ovary (CHO) cells were investigated. The two drugs studied were 1,4-dihydroxy-5,8-bis-(2-[2-hydroxyethyl)amino)ethylamino)-9,10-anthracenedione (DHAQ, NSC No. 279836) and 1,4-bis-(2-[2-hydroxyethyl)-amino)ethylamino)-9,10-anthracenedione (HAQ, NSC No. 287513). DHAQ was 100-fold more potent in reducing cell survival than HAQ. DNA strand breaks were assayed by alkaline elution. DHAQ (10 ng/ml) caused more strand breakage than 1000 ng/ml HAQ. This difference correlates well with their differences in ability to kill cells.  相似文献   

13.
Peptide nucleic acids (PNA) mimic DNA and RNA by forming complementary duplex structures following Watson-Crick base pairing. A set of reporter compounds that bind to DNA by intercalation are known, but these compounds do not intercalate in PNA/DNA hybrid duplexes. Analysis of the hybrid PNA duplexes requires development of reporter compounds that probe their chemical and physical properties. We prepared a series of anthraquinone (AQ) derivatives that are linked to internal positions of a PNA oligomer. These are the first non-nucleobase functional groups that have been incorporated into a PNA. The resulting PNA(AQ) conjugates form stable hybrids with complementary DNA oligomers. We find that when the AQ groups are covalently bound to PNA that they stabilize the hybrid duplex and are, at least partially, intercalated.  相似文献   

14.
The persistence and replication of defined circular and linear plasmid DNA molecules microinjected into fertilized eggs of Xenopus laevis were analyzed. For all plasmids tested, a small fraction of microinjected circular molecules was replicated; however, the overall copy numbers of either free form I or form II molecules usually did not increase through blastulation. In contrast, extensive amplification of input DNA sequences was seen whenever the microinjected DNA was assembled into high molecular weight concatemers. Moreover, the appearance and subsequent replication of injected sequences in high molecular weight DNA were enhanced when linear (form III), rather than circular, molecules were microinjected. The injected form III DNA was rapidly converted into long linear concatemers. All possible orientations of monomeric molecules within the concatemers were observed although, on occasion, head-to-tail orientations were favored. Long linear concatemers were replicated very efficiently, irrespective of the sequence of the input DNA. Form I and form II DNA molecules were also formed in the embryo from microinjected form III DNA. A small fraction of these circular forms was replicated, although overall copy numbers did not increase significantly. Form III molecules that remained monomeric were not observed to be replicated at all within our limits of detection. In some batches of embryos, form I and form II DNA molecules were replicated to the extent that overall copy number increased. Even in these cases, however, the amplification of long linear concatemers of the input DNA sequences was more efficient.  相似文献   

15.
Ametantrone and mitoxantrone, two new anthracenedione antineoplastic agents, produced a concentration-dependent inhibition of hepatic microsomal lipid peroxidation. Malondialdehyde production was diminished from 10.6 nmoles/mg/60 min to 3.3 and 5.4 nmoles/mg/60 min, in the presence of 100 μM mitoxantrone and ametantrone, respectively. Under similar conditions, Adriamycin stimulated lipid peroxidation over twofold. In addition, both mitoxantrone and ametantrone inhibited Adriamycin-stimulated lipid peroxidation, with 50% inhibition occurring at concentrations of 4 and 6 μM, respectively. Microsomal superoxide production was not significantly inhibited at anthracenedione concentrations which markedly decreased lipid peroxidation, suggesting that inhibition of lipid peroxidation was not the result of inhibition of superoxide generation. These results correlate with the lack of anthracenedione cardiotoxicity and also demonstrate anthracenedione inhibition of lipid peroxidation at micromolar concentrations; an observation with potential therapeutic significance.  相似文献   

16.
The photosensitizing efficiencies of eight dyes have been compared; two acridines, two xanthene derivatives, one sulphur-containing dye and three chemotherapeutic agents. The analysed reaction was the photosensitized induction of free radicals in calf-thymus DNA at low temperature. The binding of these dyes to DNA was first measured. Both strong (process I) and weak (process II) binding, with different intensities, either alone or together, were observed as mode of fixation. Whatever the nature of their binding, all the dyes used revealed a photosensitizing power as inducers of peroxide radicals in DNA. Their relative efficiencies, expressed as a function of the amount of dye molecules bound to DNA, were found to be very different. Intercalation, however, appeared to favour the free-radical induction as the first strongly bound molecules were more efficient.  相似文献   

17.
The two membrane precursors (pentapeptide lipids I and II) of peptidoglycan are present in Escherichia coli at cell copy numbers no higher than 700 and 2,000 respectively. Conditions were determined for an optimal accumulation of pentapeptide lipid II from UDP-MurNAc-pentapeptide in a cell-free system and for its isolation and purification. When UDP-MurNAc-tripeptide was used in the accumulation reaction, tripeptide lipid II was formed, and it was isolated and purified. Both lipids II were compared as substrates in the in vitro polymerization by transglycosylation assayed with PBP 1b or PBP 3. With PBP 1b, tripeptide lipid II was used as efficiently as pentapeptide lipid II. It should be stressed that the in vitro PBP 1b activity accounts for at best to 2 to 3% of the in vivo synthesis. With PBP 3, no polymerization was observed with either substrate. Furthermore, tripeptide lipid II was detected in D-cycloserine-treated cells, and its possible in vivo use in peptidoglycan formation is discussed. In particular, it is speculated that the transglycosylase activity of PBP 1b could be coupled with the transpeptidase activity of PBP 3, using mainly tripeptide lipid II as precursor.  相似文献   

18.
Several anthraquinone derivatives are active against different kinds of human cancer. The cancerostatic activity has been mainly attributed to their ability to bind strongly to DNA by intercalation. Here, infrared spectroscopy was used to detect further, more specific DNA interactions with the prominent anticancer drugs daunomycin, adriamycin, aclacinomycin A and mitoxantrone as well as with the cytotoxic violamycin BI. The most striking result was a significant decrease in wave number of the band arising from antisymmetric stretching vibration of the PO2- groups of DNA upon complexation with adriamycin, aclacinomycin A, violamycin BI and mitoxantrone. This became evident after separation of the contributions from conformational changes of DNA to the influence on the wave number of that band. The drug-induced shift was interpreted in terms of the formation of a hydrogen bond between the intercalated drug molecules and the PO2- moiety of DNA via the following terminal hydroxyl groups: C14-OH for adriamycin, C4-OH for both aclacinomycin A and violamycin BI and, more tentatively, the external side-chain OH of mitoxantrone. Theoretical considerations, consisting of semi-empirical CNDO/2 calculations as well as normal coordinate analyses performed with molecular model fragments, provided results confirming and rationalising the experimental findings. The capacities of the anthracyclines for restriction of the conformational flexibility of DNA differ, presumably due to variations in the spatial dimensions of the sugar moieties of the drugs. The compatibility of the present results with data obtained from current geometrical models, especially those for the DNA-daunomycin and DNA-adriamycin complexes, is discussed in detail.  相似文献   

19.
We examined molecular events and morphological features associated with apoptosis induced by anthraquinone anticancer drugs aclarubicin, mitoxantrone and doxorubicin in two spontaneously immortalized cell lines (NIH 3T3 and B14) in relation to cytotoxicity of these drugs. The investigated cells showed similar sensitivity to aclarubicin but different sensitivity to doxorubicin and mitoxantrone: mitoxantrone was the most cytotoxic drug in both cell lines. All three drugs triggered both apoptosis and necrosis but none of these processes was positively correlated with their cytotoxicity. Apoptosis was the prevalent form of cell kill by aclarubicin, while doxorubicin and mitoxantrone induced mainly the necrotic mode of cell death. The extent and the timing of apoptosis were strongly dependent on the cell line, the type of the drug and its dose, and were mediated by caspase-3 activation. A significant increase in caspase-3 activity and the percentage of apoptotic cells, oligonucleosomal DNA fragmentation, chromatin condensation and formation of apoptotic bodies was observed predominantly in B14 cells. NIH 3T3 cells showed lesser changes and a lack of DNA fragmentation. Aclarubicin was the fastest acting drug, inducing DNA fragmentation 12 h earlier than doxorubicin, and 24 h earlier than mitoxantrone. Caspase-3 inhibitor Ac-DEVD-CHO did not show any significant effect on drug cytotoxicity and DNA nucleosomal fragmentation.  相似文献   

20.
Previous studies with HCHO have revealed a reaction with superhelical DNA that strongly suggests that this DNA consists of small regions of interrupted secondary structure. To map these sites in PM2 DNA, the following set of experiments was performed using electron microscopy. (i) A denaturation map of nicked form II was obtained using Inman's alkaline-HCHO conditions. (ii) The superhelical form I was reacted with HCHO at 30 C until equilibrium was achieved at the interrupted sites (3.6% reactivity). The excess HCHO was removed rapidly and X-ray treatment was employed to nick these prereacted molecules. These form II molecules containing HCHO (form II HCHO) were also subjected to denaturation mapping. It would be expected that the HCHO-unpaired regions would serve as induction sites for the propagation of melting. Hence, depending on the location of the induction sites; we would anticipate either the creation of new regions of melting or a normal denaturation map shifted to lower pH values. Comparison of the development of progressive denaturation of form II and form II HCHO reveals that the latter is the case. The denaturation maps of form II are highly organized patterns of adenine-thymine (AT)-rich regions, with a total of five regions at extreme pH conditions. There are six highly organized regions for form II HCHO, i.e., smaller adjacent loops, at low denaturation conditions where no denaturation is seen for form II. These coalesce into the pattern for form II containing four of five A-T-rich regions observed for form II. Hence we conclude that the regions of altered hydrogen bonding in superhelical PM2 DNA are four to six in number and they map in the A-T-rich regions of the 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