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 375 毫秒
1.
Eukaryotic genomes are partitioned into active and inactive domains called euchromatin and heterochromatin, respectively. In Neurospora crassa, heterochromatin formation requires methylation of histone H3 at lysine 9 (H3K9) by the SET domain protein DIM-5. Heterochromatin protein 1 (HP1) reads this mark and directly recruits the DNA methyltransferase, DIM-2. An ectopic H3 gene carrying a substitution at K9 (hH3(K9L) or hH3(K9R)) causes global loss of DNA methylation in the presence of wild-type hH3 (hH3(WT)). We investigated whether other residues in the N-terminal tail of H3 are important for methylation of DNA and of H3K9. Mutations in the N-terminal tail of H3 were generated and tested for effects in vitro and in vivo, in the presence or absence of the wild-type allele. Substitutions at K4, K9, T11, G12, G13, K14, K27, S28, and K36 were lethal in the absence of a wild-type allele. In contrast, mutants bearing substitutions of R2, A7, R8, S10, A15, P16, R17, K18, and K23 were viable. The effect of substitutions on DNA methylation were variable; some were recessive and others caused a semi-dominant loss of DNA methylation. Substitutions of R2, A7, R8, S10, T11, G12, G13, K14, and P16 caused partial or complete loss of DNA methylation in vivo. Only residues R8-G12 were required for DIM-5 activity in vitro. DIM-5 activity was inhibited by dimethylation of H3K4 and by phosphorylation of H3S10, but not by acetylation of H3K14. We conclude that the H3 tail acts as an integrating platform for signals that influence DNA methylation, in part through methylation of H3K9.  相似文献   

2.
3.
The interactions of the steroidal and nonsteroidal estrogen-platinum (Pt) (II)-based anticancer drugs 16beta-hydroxymethyl-16alpha-[8-(2-pyridin-2-yl-ethylamino)-3,6-dioxaoctyl]-1,3,5(10)-estratrien-3,17betadiol dichloroplatinum (II) (JPM-39), 4-[6-(2'-pyridylethylamino)-butyloxy)-phenyl]-7-methoxy-2,2-dimethyl-3-phenyl-chroman dichloroplatinum (II) (ATG-99), and 1-[(2-aminoethyl)amino]-9,10,10-tris(4-hydroxyphenyl)-9-decene dichloroplatinum (II) (GEB-28) with calf-thymus DNA in vitro using constant DNA concentration and various drug levels were studied. Fourier transform infrared (FTIR) and circular dichroism (CD) were studied with calf-thymus DNA in vitro using constant DNA concentration and various drug levels. FTIR, UV-visible, and CD spectroscopic methods were used to characterize the drug binding mode, the binding constant, and structural variations of DNA in aqueous solution. Spectroscopic evidence showed that the various Pt-based drugs bind indirectly to the major and minor grooves of DNA duplex with some degree of drug-phosphate interaction. The overall binding constants for JPM-39, GEB-28, and ATG-99 are K(JPM-39) = 4.2 (+/-0.75) x 10(3) M(-1), K(GEB-28) = 3.4 (+/-0.65) x 10(3) M(-1), and K(ATG-99) = 2.1 (+/-0.45) x 10(3) M(-1). DNA aggregation occurs at high drug concentration, while DNA remains in the B-family structure.  相似文献   

4.
Temperature-sensitive (ts) mutants in a number of complementation groups of herpes simplex virus type 1 (HSV-1) are deficient in DNA polymerase induction at the restrictive temperature. Twenty-two mutants in 15 complementation groups were tested for sensitivity to phosphonoacetate (PAA), a compound that inhibits HSV replication in vivo and the DNA polymerase in vitro. One mutant, tsD9, was resistant to PAA (Pr), whereas all others were sensitive. Revertants of tsD9 to the ts+ phenotype simultaneously lost PAA resistance. Additional Pr mutants were isolated from ts mutants belonging to several complementation groups of HSV-1. Double mutants (ts Pr phenotype) were used in three-factor recombination analyses to locate the PAA locus on the genetic map at a position indistinguishable from the ts lesion in tsD9. In all cases, resistance or sensitivity to PAA in vivo was correlated with resistance or sensitivity of DNA polymerase in vitro. These data are compatible with the temperature-sensitive lesion of tsD9 and the determinant of PAA sensitivity both residing in the structural gene for DNA polymerase.  相似文献   

5.
The well-reported, but moderate antitumor activity of the acronycine alkaloid led us to synthesize a novel series of thioacridone compounds related to acronycine, as potential anticancer agents. Compounds were designed either as DNA intercalating agents, or as DNA intercalating agents with covalent bond forming potential. Bathochromic shifts of the compounds upon complexation with salmon testis DNA suggested intercalation as the mode of DNA binding. The binding interaction of the compounds was found to be approximately 10(2) M(-1), with that of the most potent compound 1-(2-dimethylaminoethylamino)-9(10H)-thioacridone, 10(4) M(-1). In vitro cytotoxic activity (IC50) against HL-60 cells was found to range between 3.5 and 22 microg/mL. QSAR analyses yielded a multiple linear regression equation with an r2 of 0.847 for DNA binding and an r2 of 0.575 for cytotoxicity. The physicochemical parameters used in the QSAR analyses were logP, polar surface area, and calculated molar refractivity. Docking studies were also performed to compare the binding of the most potent and least potent compounds in the study in order to predict desirable chemical characteristics for further exploitation in drug design efforts. The thioacridone compounds in this series demonstrate cytotoxic activity in vitro that merit future in vivo evaluation.  相似文献   

6.
The ability of 9-aminoacridine to induce mutagenic lesions during DNA replication in vitro was investigated. The ampicillinase gene of pBR322 was replicated in vitro in the presence of 9-aminoacridine. Transfection of the replicated DNA into Escherichia coli gave Amps mutants. Determination of the base changes in 76 of these mutants indicated that the spectrum of mutations induced by 9-aminoacridine was consistent with its action in vivo. Both large (407-base) and small (1- and 2-base) deletions were induced at repetitive sequences. The frequency of deletion mutations depended on the identity of the base deleted and sequences surrounding the deletions. The characteristics of the frameshift mutations induced were consistent with the interactions of 9-aminoacridine with DNA. These results establish that 9-aminoacridine can induce frameshift mutations during the replication process and provide an in vitro model of frameshift induction for mechanistic studies.  相似文献   

7.
Methyl-accepting assays and a sensitive method for labeling specific CpG sites have been used to show that the DNA of F9 embryonal carcinoma cells decreases in 5-methylcytosine content by ca. 9% during retinoic acid-induced differentiation, whereas the DNA of dimethyl sulfoxide-induced Friend murine erythroleukemia (MEL) cells loses ca. 3.8% of its methyl groups. These values correspond to the demethylation of 2.2 X 10(6) and 0.9 X 10(6) 5'-CpG-3' sites per haploid genome in differentiating F9 and MEL cells, respectively. Fluorography of DNA restriction fragments methylated in vitro and displayed on agarose gels showed that demethylation occurred throughout the genome. In uninduced F9 cells, the sequence TCGA tended to be more heavily methylated than did the sequence CCGG, whereas this tendency was reversed in MEL cells. The kinetics of in vitro DNA methylation reactions catalyzed by MEL cell DNA methyltransferase showed that substantial numbers of hemimethylated sites accumulate in the DNA of terminally differentiating F9 and MEL cells, implying that a partial loss of DNA-methylating activity may accompany terminal differentiation in these two cell types.  相似文献   

8.
Cytotoxic effects and topoisomerase II-mediated DNA breaks induced in vitro by ellipticine derivatives were examined in connection with 1H NMR and circular dichroism (CD) studies on molecular structures and interactions of drugs with DNA. The compounds included four 9-hydroxyellipticine and two 7-hydroxyisoellipticine derivatives. Structure-activity relationships indicated that a change in nitrogen atom position in the pyridinic ring greatly affected drug effects both on topoisomerase II action and cytotoxicity to L1210 cells. The four 9-hydroxyellipticine derivatives yielded bell-shaped curves in in vitro topoisomerase II-mediated DNA break assays, whereas the two 7-hydroxyisoellipticine derivatives demonstrated an almost linear increase at the same concentration (0-10 microM). In both cases, the intensity of cleavage was modulated by the position and the degree of methylation on the pyridinic ring, and results were correlated with cytotoxic activity expressed as the in vitro ID50 values for L1210 leukemia cells. 1H NMR experiments performed on free drug molecules in solution revealed that the two protons (alpha and beta) contiguous to the biologically important hydroxy group were sensitive to changes in electron distribution produced by the distant chemical modifications and methylations of the pyridinic ring. A linear relationship was observed between the differences in chemical shifts of alpha and beta protons (delta delta alpha-beta) versus ID50 values. CD experiments indicated that, at weak ionic strength I = 0.02 and at pH 7, drugs interact with the poly[d(A-T)] duplex according to a "three-mode binding model" which is governed by the drug structure and the drug to DNA ratio. The intercalation mode was related to the induction of topoisomerase II-mediated DNA cleavage, while the external binding mode consecutive to intercalation was related to cleavage suppression. These two modes concerned the good intercalators 9-hydroxyellipticines. The third was found for the weak intercalators 7-hydroxyisoellipticines and was characterized by self-stacked molecules bound "outside" DNA, presumably in the minor groove. Ligands either could be intercalated partially or linked at the edge of bases with a small number of molecules filling intercalation sites, for the second alternative. In addition to having different binding modes, 9-hydroxyellipticines were better inducers of DNA distortions than 7-hydroxyisoellipticines. The incidence of the drug binding modes on DNA-topoisomerase II recognition was discussed in connection with the in vitro cytotoxic activity exhibited by the drugs.  相似文献   

9.
10.
An SV40-based in vitro replication system has been used to examine the effects of platinum compounds on eukaryotic DNA replication. Plasmid templates containing the SV40 origin of replication were modified with the anticancer drug cis-diamminedichloroplatinum(II) (cis-DDP, cisplatin) or the inactive analogues [Pt(dien)Cl]+ and trans-DDP. The platinated plasmids were used as templates for DNA synthesis by the DNA polymerases present in cytosolic extracts prepared from human cell lines HeLa and 293. Bifunctional adducts formed by cis- and trans-DDP inhibited DNA replication by 95% at a bound drug to nucleotide ratio [(D/N)b] of less than 9 x 10(-4), in contrast to the monofunctional [Pt(dien)Cl]+ analogues, which required a (D/N)b of 3.4 x 10(-3) for 62% inhibition of DNA replication. An average of two platinum adducts per genome was sufficient for inhibition of DNA replication by cisplatin. When trans-DDP-modified, but not cis-DDP-modified, SV40 origin containing plasmids [(D/N)b = 1.7 x 10(-3)] were allowed to incubate in the 293 cytosolic extracts for 1 h prior to addition of T-antigen to initiate replication, DNA synthesis was restored to 30% of control. This result suggested the presence of an activity in the extracts that reactivates trans-DDP-modified DNA templates for replication. This hypothesis was confirmed by an in vitro nucleotide excision repair assay that revealed activity in 293 and HeLa cell extracts selective for trans-DDP-modified plasmid DNAs. Such selective repair of trans-DDP-damaged DNA in human cells would contribute to its lack of antitumor activity.  相似文献   

11.
A marker rescue system based on the repair of the kanamycin resistance gene nptII was constructed for use in Gram-positive bacteria and established in Bacillus subtilis 168. Marker rescue was detected in vitro using different types of donor DNA containing intact nptII. The efficiency of marker rescue using chromosomal DNA of E. coli Sure as well as plasmids pMR2 or pSR8-30 ranged from 3.8 x 10(-8) to 1.5 x 10(-9) transformants per nptII gene. Low efficiencies of ca. 10(-12) were obtained with PCR fragments of 792 bp obtained from chromosomal DNA of E. coli Sure or DNA from a transgenic potato. B. subtilis developed competence during growth in milk and chocolate milk, and marker rescue transformation was detected with frequencies of ca. 10(-6) and 10(-8), respectively, using chromosomal DNA of E. coli Sure as donor DNA. Although the copy number of nptII genes of the plant DNA exceeded that of chromosomal E. coli DNA in the marker rescue experiments, a transfer of DNA from the transgenic plant to B. subtilis was detectable neither in vitro nor in situ.  相似文献   

12.
A comparison between [14C]aflatoxin B1 (AFB1) and [14C]aflatoxin G1 (AFG1) binding to rat liver and kidney cellular macromolecules has shown AFG1-DNA and-ribosomal RNA binding to be lower in both organs. For both mycotoxins more was bound to nucleic acids than to protein. Two hours after intraperitoneal injection (60 microgram/100 g) of [14C] AFB1, 40 ng, 151 ng/mg. Loss of radioactivity bound to liver DNA for both [14C]AFB1 and protein respectively and for [14C]AFG1 the respective figures were 10, 7 and 1 ng/mg. Loss of liver bound radioactivity to DNA for both [14C]AFG1 and [14C]AFG1 appeared to be biphasic indicating that an enzymic DNA repair process may be operating. In vitro binding studies also showed less AFG1 was bound to exogenous DNA after microsomal activation than AFB1. This difference was not a result of differences in the chemical reactivity of the "ultimate" electrophilic species, the respective expoxides, since chemical activation studies using 3-chloroperbenzoic acid showed similar amounts of AFG1 and AFB1 to be converted to the epoxides and to bind to DNA. Studies on the distribution coefficients of the two mycotoxins showed AFB1 to be more lipophilic than AFG1 and this may be an important factor in determining the weaker carcinogenicity of the latter compound. Characterisation of the major AFG1-DNA adduct formed in vitro, in vivo and after peracid oxidation showed it to have the structure trans-9,10-dihydro-9-(7-guanyl)-10-hydroxy-aflatoxin G1. This adduct is similar to that obtained from AFB1 by activation in vivo, in vitro and after peracid oxidation.  相似文献   

13.
DNA-damaging activity of patulin in Escherichia coli   总被引:1,自引:0,他引:1  
At a concentration of 10 micrograms/ml, patulin caused single-strand DNA breaks in living cells of Escherichia coli. At 50 micrograms/ml, double-strand breaks were observed also. Single-strand breaks were repaired in the presence of 10 micrograms of patulin per ml within 90 min when the cells were incubated at 37 degrees C in M9-salts solution without a carbon source. The same concentration also induced temperature-sensitive lambda prophage and a prophage of Bacillus megaterium. When an in vitro system with permeabilized Escherichia coli cells was used, patulin at 10 micrograms/ml induced DNA repair synthesis and inhibited DNA replication. The in vivo occurrence of DNA strand breaks and DNA repair correlated with the in vitro induction of repair synthesis. In vitro the RNA synthesis was less affected, and overall protein synthesis was not inhibited at 10 micrograms/ml. Only at higher concentrations (250 to 500 micrograms/ml) was inhibition of in vitro protein synthesis observed. Thus, patulin must be regarded as a mycotoxin with selective DNA-damaging activity.  相似文献   

14.
The feasibility of using probes directed towards ribosomal DNAs (rDNAs) as a quantitative approach to estimating cell numbers was examined and applied to study the structure of a bacterial community in humic acid-rich salt marsh sediments. Hybridizations were performed with membrane-bound nucleic acids by using seven group-specific DNA oligonucleotide probes complementary to 16S rRNA coding regions. These included a general eubacterial probe and probes encompassing most members of the gram-negative, mesophilic sulfate-reducing bacteria (SRB). DNA was extracted from sediment samples, and contaminating materials were removed by a series of steps. Efficiency of DNA extraction was 48% based on the recovery of tritiated plasmid DNA added to samples prior to extraction. Reproducibility of the extraction procedure was demonstrated by hybridizations to replicate samples. Numbers of target cells in samples were estimated by comparing the amount of hybridization to extracted DNA obtained with each probe to that obtained with a standard curve of genomic DNA for reference strains included on the same membrane. In June, numbers of SRB detected with an SRB-specific probe ranged from 6.0 x 10(7) to 2.5 x 10(9) (average, 1.1 x 10(9) +/- 5.2 x 10(8)) cells g of sediment-1. In September, numbers of SRB detected ranged from 5.4 x 10(8) to 7.3 x 10(9) (average, 2.5 x 10(9) +/- 1.5 x 10(9)) cells g of sediment-1. The capability of using rDNA probes to estimate cell numbers by hybridization to DNA extracted from complex matrices permits initiation of detailed studies on community composition and changes in communities based on cell numbers in formerly intractable environments.  相似文献   

15.
At a concentration of 10 micrograms/ml, patulin caused single-strand DNA breaks in living cells of Escherichia coli. At 50 micrograms/ml, double-strand breaks were observed also. Single-strand breaks were repaired in the presence of 10 micrograms of patulin per ml within 90 min when the cells were incubated at 37 degrees C in M9-salts solution without a carbon source. The same concentration also induced temperature-sensitive lambda prophage and a prophage of Bacillus megaterium. When an in vitro system with permeabilized Escherichia coli cells was used, patulin at 10 micrograms/ml induced DNA repair synthesis and inhibited DNA replication. The in vivo occurrence of DNA strand breaks and DNA repair correlated with the in vitro induction of repair synthesis. In vitro the RNA synthesis was less affected, and overall protein synthesis was not inhibited at 10 micrograms/ml. Only at higher concentrations (250 to 500 micrograms/ml) was inhibition of in vitro protein synthesis observed. Thus, patulin must be regarded as a mycotoxin with selective DNA-damaging activity.  相似文献   

16.
17.
The human DNA repair protein XRCC1 was overexpressed as a histidine-tagged polypeptide (denoted XRCC1-His) in Escherichia coli and purified in milligram quantities by affinity chromatography. XRCC1-His complemented the mutant Chinese hamster ovary cell line EM9 when constitutively expressed from a plasmid or when introduced by electroporation. XRCC1-His directly interacted with human DNA ligase III in vitro to form a complex that was resistant to 2 M NaCl. XRCC1-His interacted equally well with DNA ligase III from Bloom syndrome, HeLa and MRC5 cells, indicating that Bloom syndrome DNA ligase III is normal in this respect. Detection of DNA ligase III on far Western blots by radiolabelled XRCC1-His indicated that the level of the DNA ligase polypeptide was reduced approximately 4-fold in the mutant EM9 and also in EM-C11, a second member of the XRCC1 complementation group. Decreased levels of polypeptide thus account for most of the approximately 6-fold reduced DNA ligase III activity observed previously in EM9. Immunodetection of XRCC1 on Western blots revealed that the level of this polypeptide was also decreased in EM9 and EM-C11 (> 10-fold), indicating that the XRCC1-DNA ligase III complex is much reduced in the two CHO mutants.  相似文献   

18.
The ability of the common aquatic contaminant 1-nitropyrene to form DNA adducts in fish was investigated in vitro and in vivo using Brown trout (Salmo trutta) and turbot (Scophthalmus maximus)in comparison to the Wistar rat. In vitro studies used Brown trout (control and induced (50 mg kg-1-naphthoflavone (NF), i.p. 3 day pre-treatment single injection)) and induced rat (PB; 0 1% w/v for 7 days in drinking water, NF; 80 mg kg-1, single injection 2 days prior to sacrifice). Hepatic 9000 g supernatant (S9 fractions) were incubated for 2 hours (at 25 C for fish and 37 C for rat) with calf thymus DNA (1mg) and 1-NP (100 M). With all S9 fractions the presence of three distinct 1-NP-related DNA adducts was detected using the butanol enrichment procedure of the 32Ppostlabelling assay. A greater level of DNA adducts was observed with the uninduced compared to the induced trout S9 (37, 12 and 8 fold greaterfor adducts in chromatograph areas 1-3 respectively) suggesting the enhancement of detoxification pathways with respect to bulky adducts following NF pre-treatment. DNA adduct levels in the induced rat consistently demonstrated approximately two-fold higher levels as compared to the induced fish, reflecting the lower protein levels in the S9 fraction of Brown trout (42 and 22 mg ml-1 for rat and fish respectively). Turbot, rat and Brown trout (uninduced and induced (NF; 50 mg kg-1; i.p. single injection 3 days prior)) were dosed with 100 mg kg-1 1-NP (i.p. single injection, 24 hours). Liver DNA from both turbot and rat exhibited a 1- NP related adduct spot which was similar in position to that of area 1 in the incubations with S9 from rat and Brown trout. However, in contrast to the in vitro studies no 1-NPrelated adducts were found in liver DNA from induced and uninduced Brown trout. This study highlights the potential, in a marine and a freshwater fish, for 1-NP metabolism to reactive of binding to DNA. However, activation of 1-NP was more optimal in the S9-mediated system, possibly reflecting the influence of detoxification systems.  相似文献   

19.
To identify viral myc proteins, we have prepared myc-specific antibodies: (i) against a synthetic peptide corresponding to the nine carboxy-terminal amino acids of the viral myc (C9); (ii) against a bacterially expressed viral myc protein obtained by inserting the SalI-BamHI fragment of the viral MC29 DNA clone in the expression vector pPLc24. Both antisera recognize a protein of 55 000 mol. wt., p55v-myc, in MH2- and OK10-transformed fibroblasts. The protein is located in the nucleus, as shown by indirect immunofluorescence and cell fractionation. Antibodies against the C9 peptide were used to purify the p55v-myc by immunoaffinity column purification (3000-fold) from OK10- and MH2-transformed fibroblasts. p55v-myc binds to double-stranded DNA in vitro as does p110gag-myc. DNA binding in vitro is inhibited by the immunoglobulin fraction of antibodies against the bacterially expressed myc protein. Furthermore, a synthetic peptide consisting of 16 amino acids (C16) was used to isolate specific immunoglobulins which also inhibit DNA binding in vitro. OK10 codes, in addition to p55v-myc, for a p200gag-pol-myc polyprotein. The majority of this protein is located in the cytoplasm (79%). The purified protein binds to single-stranded RNA in vitro, unlike other gag-myc or myc proteins.  相似文献   

20.
Xiao Y  Chandra M  Silverman SK 《Biochemistry》2010,49(44):9630-9637
We recently reported the identification by in vitro selection of 10MD5, a deoxyribozyme that requires both Mn2+ and Zn2+ to hydrolyze a single-stranded DNA substrate with formation of 5′-phosphate and 3′-hydroxyl termini. DNA cleavage by 10MD5 proceeds with kobs=2.7 h(?1) and rate enhancement of 10(12) over the uncatalyzed P?O hydrolysis reaction. 10MD5 has a very sharp pH optimum near 7.5, with greatly reduced DNA cleavage rate and yield when the pH is changed by only 0.1 unit in either direction. Here we have optimized 10MD5 by reselection (in vitro evolution), leading to variants with broader pH tolerance, which is important for practical DNA cleavage applications. Because of the extensive Watson?Crick complementarity between deoxyribozyme and substrate, the parent 10MD5 is inherently sequence-specific; i.e., it is able to cleave one DNA substrate sequence in preference to other sequences. 10MD5 is also site-specific because only one phosphodiester bond within the DNA substrate is cleaved, although here we show that intentionally creating Watson?Crick mismatches near the cleavage site relaxes the site specificity. Newly evolved 10MD5 variants such as 9NL27 are also sequence-specific. However, the 9NL27 site specificity is relaxed for some substrate sequences even when full Watson?Crick complementarity is maintained, corresponding to a functional compromise between pH tolerance and site specificity. The site specificity of 9NL27 may be restored by expanding its “recognition site” from ATGT (as for 10MD5) to ATGTT or larger, i.e., by considering 9NL27 to have reduced substrate sequence tolerance relative to 10MD5. These findings provide fundamental insights into the interplay among key deoxyribozyme characteristics of tolerance and selectivity, with implications for ongoing development of practical DNA-catalyzed DNA hydrolysis.  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