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.
Asymmetric Transcription of Bacteriophage Mu-1   总被引:8,自引:2,他引:6  
The deoxyribonucleic acid (DNA) of bacteriophage Mu-1 can be separated into its complementary strands by poly(U,G) binding and equilibrium centrifugation. DNA-ribonucleic acid (RNA) hybridizations in liquid show that more than 98% of "early" phage-specific RNA and over 96% of "late" messenger species bind to the heavy [poly(U,G)-binding] strand of Mu-1 DNA. A small (1.45%) but significant amount of late RNA binds to the light strand. The significance of this RNA fraction is discussed in connection with the peculiar structure of denatured and reannealed Mu-1 DNA.  相似文献   

2.
Although the functional target of quinolone antibacterials such as nalidixic acid and norfloxacin has been identified as the enzyme DNA gyrase, the direct binding site of the drug is the DNA molecule [Shen, L. L., & Pernet, A. G. (1985) Proc. Natl. Acad. Sci. U.S.A. 82, 307-311]. As described in this paper, binding specificity and cooperativity of quinolones to DNA were further investigated with the use of a variety of DNA species of different structures and different base compositions. Results show that the drug binding specificity is controlled and determined largely by the DNA structure. The drug binds weakly and demonstrates no base preference when DNA strands are paired. The drug binds with much greater affinity when the strands are separated, and consequently, binding preference emerges: it binds better to poly(G) and poly(dG) over their counterparts including poly(dI). The results suggest that the drug binds to unpaired bases via hydrogen bonding and not via ring stacking with DNA bases. The weak binding to relaxed double-stranded DNA and the stronger binding to single-stranded DNA are both nonspecific as they do not demonstrate binding saturation and cooperativity. The specific type of binding, initially demonstrated in our previous publication with the supercoiled DNA and more recently with complex formed between linear DNA and DNA gyrase [Shen, L. L., Kohlbrenner, W. E., Weigl, D., & Baranowski, J. (1989) J. Biol. Chem. (in press)], occurs near the drug's supercoiling inhibition concentration. As shown in this paper, binding saturation curves of this type are highly cooperative (with Hill constant greater than 4).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
The complementary strands of adenovirus type 2 (Ad2) DNA were separated by buoyant density gradient centrifugation with poly (U, G). The complementary strand DNA was shown to remain intact through the course of strand separation. The l-strand of Ad2 DNA, appearing in the less dense complex with poly (U, G) in neutral CsCl density gradients, was shown to have a buoyant density in alkaline (pH 12.5) CsCl density gradients which is 2 to 3 mg per ml greater than that of its complement (h-strand). Renaturation of purified complementary strand DNA was observed only in mixtures of h- and l-strand DNA, and then with the second-order reaction rate expected for Ad2 DNA. Hybridization of the complementary strands of Ad2 DNA with cytoplasmic mRNA isolated from infected HeLa cells was performed in liquid phase and analyzed by hydroxylapatite chromatography. Before viral DNA synthesis (6 h after infection), 13 to 18% of the h-strand and 30 to 35% of the l-strand were represented in viral mRNA. Late (18 h) after infection the mRNA represented 20 to 25% and 63 to 68% of the h- and l-strands, respectively. Most, if not all sequences present in viral mRNA before viral DNA synthesis were also present in the cytoplasm late in infection.  相似文献   

4.
The interaction of the nuclear estrogen receptor from hen oviduct with nucleic acids were studied by competition assay using DNa-cellulose centrifugation. We demonstrated that the estradiol-receptor complex binds similarly well to poly(A) RNA and denatured DNA. The estrogen receptor was found to interact more strongly with poly(G), poly(U) than with poly(A), poly(C). The receptor complex binds similarly to poly(A) and poly(dA), and to poly(U) and poly(dU). However, the receptor complex shows stronger binding to poly(G) than to poly(dG) and to poly(C) than to poly(dC). Studies with heteropolyribonucleotides indicated that poly(U1G1) is more effective in competing for the estrogen receptor, and poly(AC) and poly(AUG) are moderately effective, whereas poly(ACU) is least effective. GMP and dGMP showed some competition for the nuclear receptor at 300-fold higher nucleotide concentrations than that of the synthetic poly(G). Observations that the nuclear estrogen receptor binds to poly(A) RNA and interacts selectively with polyribonucleotides suggest that the estrogen receptor-RNA interaction may play a role for the function of estrogens in gene regulation.  相似文献   

5.
A Ray  G S Kumar  S Das  M Maiti 《Biochemistry》1999,38(19):6239-6247
The interaction of aristololactam-beta-D-glucoside (ADG), a DNA intercalating alkaloid, with the DNA triplexes, poly(dT). poly(dA)xpoly(dT) and poly(dC).poly(dG)xpoly(dC+), and the RNA triplex poly(rU).poly(rA)xpoly(rU) was investigated by circular dichroic, UV melting profile, spectrophotometric, and spectrofluorimetric techniques. Comparative interaction with the corresponding Watson-Crick duplexes has also been examined under identical experimental conditions. Triplex formation has been confirmed from biphasic thermal melting profiles and analysis of temperature-dependent circular dichroic measurements. The binding of ADG to triplexes and duplexes is characterized by the typical hypochromic and bathochromic effects in the absorption spectrum, quenching of steady-state fluorescence intensity, a decrease in fluorescence quantum yield, an increase or decrease of thermal melting temperatures, and perturbation in the circular dichroic spectrum. Scatchard analysis indicates that ADG binds both to the triplexes and the duplexes in a noncooperative manner. Binding parameters obtained from spectrophotometric measurements are best fit by the neighbor exclusion model. The binding affinity of ADG to the DNA triplexes is substantially stronger than to the RNA triplex. Thermal melting study further indicates that ADG stabilizes the Hoogsteen base-paired third strand of the DNA triplexes whereas it destabilizes the same strand of RNA triplex but stabilizes its Watson-Crick strands. Comparative data reveal that ADG exhibits a stronger binding to the triple helical structures than to the respective double helical structures.  相似文献   

6.
Covalent circular sex-factor DNA has been isolated from donor and recipient cells during the conjugation of normal and temperature-sensitive DNA mutants of Escherichia coli. Single strands of sex-factor DNA were centrifuged in cesium chloride-poly(U,G) gradients to give two components that have been identified by annealing experiments as the separated complementary strands. When matings are performed with either DNA temperature-sensitive donor or recipient cells, the inhibition of vegetative DNA synthesis at the restrictive temperature does not interfere with transfer and circularization of the sex-factor DNA. If DNA is isolated from temperature-sensitive donor cells mated at the restrictive temperature, a specific stimulation of sex-factor DNA synthesis can be demonstrated. By separating the complementary strands of the sex-factor in a cesium chloride-poly (U,G) gradient, this DNA synthesis has been found to be asymmetric. The sex-factor DNA strand which is synthesized in the donor has the same polarity as the strand which is transferred to the recipient.  相似文献   

7.
The triplexes formed by pyrimidine alpha-oligodeoxynucleotides, 15mers alpha dT(15) or 12mers alpha dCT having dimethoxyethyl (PNHdiME), morpholino (PMOR) or propyl (PNHPr) non-ionic phosphoramidate linkages with DNA duplex targets have been investigated by UV and FTIR spectroscopy. Due to the decrease in the electrostatic repulsion between partner strands of identical lengths all modifications result in triplexes more stable than those formed with unmodified phosphodiester beta-oligodeoxynucleotides (beta-ODNs). Among the alpha-ODN third strands having C and T bases and non-ionic phosphoramidate linkages (alpha dCTPN) the most efficient modification is (PNHdiME). The enhanced third strand stability of the alpha dCTPN obtained as diastereoisomeric mixtures is attenuated by the steric hindrance of the PMOR linkages or by the hydrophobicity of the PNHPr linkages. All alpha dCTPN strands form triplexes even at neutral pH. In the most favorable case (PNHdiME), we show by FTIR spectroscopy that the triplex formed at pH 7 is held by Hoogsteen T*A.T triplets and in addition by an hydrogen bond between O6 of G and C of the third strand (Tm = 30 degrees C). The detection of protonated cytosines is correlated at pH 6 with a high stabilization of the triplex (Tm = 65 degrees C). While unfavorable steric effects are overcome with alpha anomers, the limitation of the pH dependence is not completely suppressed. Different triplexes are evidenced for non pH dependent phosphoramidate alpha-thymidilate strands (alpha dT(15)PN) interacting with a target duplex of identical length. At low ionic strength and DNA concentration we observe the binding to beta dA(15) either of alpha dT(15)PN as duplex strand and beta dT(15) as third strand, or of two hydrophobic alpha dT(15)PNHPr strands. An increase in the DNA and counterion concentration stabilizes the anionic target duplex and then the alpha dT(15)PN binds as Hoogsteen third strand.  相似文献   

8.
The carcinogen 1-methyl-3-hydroxyxanthine after esterification binds covalently to polynucleotides, RNA and DNA. All four ribopolynucleotides and poly(dT) are targets. Depending on reaction conditions, covalent binding is greatest to poly(A) followed by poly(U), poly(dT), poly(G), poly(C), RNA and DNA. Maximal covalent modification of DNA is one moiety per 360 nucleotides. All modified polynucleotides, RNA and DNA, except poly guanylic acid have been enzymatically digested and the major adducts characterized as nucleosides.  相似文献   

9.
UV-absorption spectrophotometry and molecular modeling have been used to study the influence of the chemical nature of sugars (ribose or deoxyribose) on triple helix stability. For the Pyrimidine.purine* Pyrimidine motif, all eight combinations were tested with each of the three strands composed of either DNA or RNA. The chemical nature of sugars has a dramatic influence on triple helix stability. For each double helix composition, a more stable triple helix was formed when the third strand was RNA rather than DNA. No stable triple helix was detected when the polypurine sequence was made of RNA with a third strand made of DNA. Energy minimization studies using the JUMNA program suggested that interactions between the 2'-hydroxyl group of the third strand and the phosphates of the polypurine strand play an important role in determining the relative stabilities of triple-helical structures in which the polypyrimidine third strand is oriented parallel to the polypurine sequence. These interactions are not allowed when the third strand adopts an antiparallel orientation with respect to the target polypurine sequence, as observed when the third strand contains G and A or G and T/U. We show by footprinting and gel retardation experiments that an oligoribonucleotide containing G and A or G and U fails to bind double helical DNA, while the corresponding DNA oligomers form stable triple-helical complexes.  相似文献   

10.
Binding of eucaryotic elongation factor Tu to nucleic acids   总被引:2,自引:0,他引:2  
The binding of eucaryotic elongation factor Tu (eEF-Tu) to nucleic acids was investigated. eEF-Tu binds to a variety of different nucleic acids with high affinity, showing a strong preference for 18 S and 28 S rRNA over transfer RNA and for ribose-containing polymers over polydeoxyribonucleotides. The factor binds at multiple sites on 28 S rRNA without strong cooperativity. eEF-Tu binds strongly to poly(G) and poly(U) but weakly, if at all, to poly(A) and poly(C). Experiments employing an airfuge demonstrate that eEF-Tu can form a quaternary complex containing the factor, 28 S rRNA, aminoacyl-tRNA, and GTP. The existence of two distinct RNA binding sites on eEF-Tu suggests that rRNA may play a role in the recognition of eEF-Tu.aminoacyl-tRNA.GTP complexes by polysomes. Support for this suggestion comes from experiments which show that poly(G) inhibits the factor-dependent binding of aminoacyl-tRNA to mRNA-programmed 80 S ribosomes. In addition, it is shown that eEF-Tu possesses an intrinsic GTPase activity which is stimulated significantly by 28 S rRNA, poly(G), and poly(U). The binding of eEF-Tu to poly(G) lowers the activation energy for eEF-Tu GTPase from 74.3 to 65.9 kJ . mol-1 and approximately doubles the Vmax of the enzymatic reaction. The results are discussed in relation to the binding of eEF-Tu to ribosomes during protein synthesis.  相似文献   

11.
Repair of thymine.guanine (T.G) and uracil.guanine (U.G) mismatched base-pairs in bacteriophage M13mp18 replicative form (RF) DNA was compared upon transfection into repair-proficient or repair-deficient Escherichia coli strains. Oligonucleotide-directed mutagenesis was used to prepare covalently closed circular heteroduplexes that contained the mismatched base-pair at a restriction recognition site. The heteroduplexes were unmethylated at dam (5'-GATC-3') sites to avoid methylation-directed biasing of repair. In an E. coli host containing uracil-DNA glycosylase (ung+), about 97% of the transfecting U.G-containing heteroduplexes had the U residue excised by the uracil-excision repair system. With the analogous T.G mispair, mismatch repair operated on almost all of the transfecting heteroduplexes and removed the T residue in about 75% of them when the mismatched T was on the minus strand of the RF DNA. Similar preferential excision of the minus-strand's mismatched base was observed whether the heteroduplex RF DNA molecules had only one or both strands unmethylated at dcm (5'-CC(A/T)GG-3') sites and whether the RF DNA was prepared by primer extension in vitro or by reannealing mutant and non-mutant DNA strands. Also, the extent and directionality of repair was the same at a U.G mispair in ung- host cells as at the analogous T.G mispair in ung- or ung+ cells. Only in a mismatch repair-deficient (mutH-) host was the plus strand of the transfecting M13mp18 heteroduplex DNA preferentially repaired. It is suggested that the plus strand nick made by the M13-encoded gene II protein might be employed by a mutH- host to initiate repair on that strand.  相似文献   

12.
Separation of the complementary strands of adenovirus type 2 DNA by poly(U,G)-CsCl density gradient centrifugation permitted studies of Ad23 DNA renaturation with independently variable concentrations of each complementary strand. Single-stranded DNA was isolated by hydroxylapatite chromatography following exhaustive incubation under such conditions, and was found to selectively represent sequences of the complement present in excess during the incubation. This result was exploited in a general method for isolation of complementary strand-specific sequences of radioactively labeled Ad2 DNA or restriction enzyme fragments of Ad2 DNA. Liquid phase saturation-hybridization experiments were carried out with labeled DNA representing each complementary strand of the six endo R.EcoRI cleavage fragments of Ad2 DNA and unlabeled messenger RNA prepared from HeLa cells late after productive infections with Ad2. The results were combined with the known endo R.EcoRI cleavage map of Ad2 DNA to construct a low-resolution map showing physically separated regions, on both complementary strands of Ad2 DNA, represented in mRNA late after infection.  相似文献   

13.
In this paper, we use molecular dynamics simulations to study the assembly of DNA-grafted nanoparticles to demonstrate specifically the effect of bidispersity in grafted DNA strand length on the thermodynamics and structure of nanoparticle assembly at varying number of grafted single-stranded DNA (ssDNA) strands and number of guanine/cytosine (G/C) bases per strand. At constant number of grafted ssDNA strands and G/C nucleotides per strand, as bidispersity in strand lengths increases, the number of nanoparticles that assemble as well as the number of neighbours per particle in the assembled cluster increases. When the number of G/C nucleotides per strand in short and long strands is equal, the long strands hybridise with the other long strands with higher frequency than the short strands hybridise with short/long strands. This dominance of the long strands leads to bidisperse systems having similar thermodynamics to that in corresponding systems with monodisperse long strands. Structurally, however, as a result of long–long, long–short and short–short strand hybridisation, bidispersity in DNA strand length leads to a broader inter-particle distance distribution within the assembled cluster than seen in systems with monodisperse short or monodisperse long strands. The effect of increasing the number of G/C bases per strand or increasing the number of grafted DNA strands on the thermodynamics of assembly is similar for bidisperse and monodisperse systems. The effect of increasing the number of grafted ssDNA strands on the structure of the assembled cluster is dependent on the extent of strand bidispersity because the presence of significantly shorter ssDNA strands among long ssDNA strands reduces the crowding among the strands at high grafting density. This relief in crowding leads to larger number of strands hybridised and as a result larger coordination number in the assembled cluster in systems with high bidispersity in strands than in corresponding monodisperse or low bidispersity systems.  相似文献   

14.
A comparative study on the interaction of sanguinarine and berberine with DNA and RNA triplexes and their parent duplexes was performed, by using a combination of spectrophotometric, UV thermal melting, circular dichroic and thermodynamic techniques. Formation of the DNA and RNA triplexes was confirmed from UV-melting and circular dichroic measurements. The interaction process was characterized by increase of thermal melting temperature, perturbation in circular dichroic spectrum and the typical hypochromic and bathochromic effects in the absorption spectrum. Scatchard analysis indicated that both the alkaloids bound to the triplex and duplex structures in a non-cooperative manner and the binding was stronger to triplexes than to parent duplexes. Thermal melting studies further indicated that sanguinarine stabilized the Hoogsteen base paired third strand of both DNA and RNA triplexes more tightly compared to their Watson-Crick strands, while berberine stabilized the third strand only without affecting the Watson-Crick strand. However, sanguinarine stabilized the parent duplexes while no stabilization was observed with berberine under identical conditions. Circular dichroic studies were also consistent with the observation that perturbations of DNA and RNA triplexes were more compared to their parent duplexes in presence of the alkaloids. Thermodynamic data revealed that binding of sanguinarine and berberine to triplexes (T.AxT and U.AxU) and duplexes (A.T and A.U) showed negative enthalpy changes and positive entropy changes but that of sanguinarine to C.GxC(+) triplex and G.C duplex exhibited negative enthalpy and negative entropy changes. Taken together, these results suggest that both sanguinarine and berberine can bind and stabilize the DNA and RNA triplexes more strongly than their respective parent duplexes.  相似文献   

15.
Native and denatured calf thymus DNA, and homopolyribonucleotides were compared with respect to chromium and protein binding after an in vitro incubation with rat liver microsomes, NADPH, and chromium(VI) or chromium(III). A significant amount of chromium bound to DNA when chromium(VI) was incubated with the native or the denatured form of DNA in the presence of microsomes and NADPH. For both native and denatured DNA the amount of protein bound to DNA increased with the amount of chromium bound to DNA. Denatured DNA had much higher amounts of chromium and protein bound than native DNA. There was no interaction between chromium(VI) and either form of DNA in the absence of the complete microsomal reducing system. The binding of chrornium(III) to native or denatured DNA was small and relatively unaffected by the presence of microsomes and NADPH. The binding of chromium and protein to polyriboadenylic acid (poly(A)), polyribocytidylic acid (poly(C), polyri-boguanylic acid (poly(G)) and polyribouridylic acid (poly(U)) was determined after incubation with chromium(VI) in the presence of microsomes and NADPH. The magnitude of chromium and protein binding to the ribo-polymers was found to be poly(G) ? poly(A) ? poly(C) ? poly(U). These results suggest that the metabolism of chromium(VI) is necessary in order for chromium to interact significantly with nucleic acids. The metabolically-produced chromium preferentially binds to the base guanine and results in DNA-protein cross-links. These findings are discussed with respect to the proposed scheme for the carcinogenicity of chromium(VI). Keywords: DNA-protein cross-links — Chromium-guanine interaction-Microsomal reduction of chromate  相似文献   

16.
The precursor terminal protein pTP is the primer for the initiation of adenovirus (Ad) DNA replication and forms a heterodimer with Ad DNA polymerase (pol). Pol can couple dCTP to pTP directed by the fourth nucleotide of the viral genome template strand in the absence of other replication proteins, which suggests that pTP/pol binding destabilizes the origin or stabilizes an unwound state. We analyzed the contribution of pTP to pTP/pol origin binding using various DNA oligonucleotides. We show that two pTP molecules bind cooperatively to short DNA duplexes, while longer DNA fragments are bound by single pTP molecules as well. Cooperative binding to short duplexes is DNA sequence independent and most likely mediated by protein/protein contacts. Furthermore, we observed that pTP binds single-stranded (ss)DNA with a minimal length of approximately 35 nt and that random ssDNA competed 25-fold more efficiently than random duplex DNA for origin binding by pTP. Remarkably, short DNA fragments with two opposing single strands supported monomeric pTP binding. pTP did not stimulate, but inhibited strand displacement by the Ad DNA binding and unwinding protein DBP. These observations suggest a mechanism in which the ssDNA affinity of pTP stabilizes Ad pol on partially unwound origin DNA.  相似文献   

17.
18.
The subject of four-stranded nucleic acid structures is reviewed. Studies on gels formed by guanosine and its analogues have provided appropriate models for the structures of poly(I) and poly(G). The stabilizing influence of certain cations, in particular K+, on Guo-5'-P gels and poly(I) is discussed in the light of recent data on selective K+ stabilization of telomeric DNA structures. The topological possibilities these dG containing sequences could adopt are discussed. In particular the role of the glycosidic linkage (anti/syn), the polarity of the strands and the orientation of the G-tetrad stacks is highlighted.  相似文献   

19.
The decoding of UGA as a selenocysteine (Sec) codon in mammalian selenoprotein mRNAs requires a selenocysteine insertion sequence (SECIS) element in the 3' untranslated region. The SECIS is a hairpin structure that contains a non-Watson-Crick base-pair quartet with a conserved G.A/A.G tandem in the core of the upper helix. Another essential component of the Sec insertion machinery is SECIS-binding protein 2 (SBP2). In this study, we define the binding site of SBP2 on six different SECIS RNAs using enzymatic and hydroxyl radical footprinting, gel mobility shift analysis, and phosphate-ethylation binding interference. We show that SBP2 binds to a variety of mammalian SECIS elements with similar affinity and that the SBP2 binding site is conserved across species. Based on footprinting studies, SBP2 protects the proximal part of the hairpin and both strands of the lower half of the upper helix that contains the non-Watson-Crick base pair quartet. Gel mobility shift assays showed that the G.A/A.G tandem and internal loop are critical for the binding of SBP2. Modification of phosphates by ethylnitrosourea along both strands of the non-Watson-Crick base pair quartet, on the 5' strand of the lower helix and part of the 5' strand of the internal loop, prevented binding of SBP2. We propose a model in which SBP2 covers the central part of the SECIS RNA, binding to the non-Watson-Crick base pair quartet and to the 5' strands of the lower helix and internal loop. Our results suggest that the affinity of SBP2 for different SECIS elements is not responsible for the hierarchy of selenoprotein expression that is observed in vivo.  相似文献   

20.
Using two direct methods we have studied the binding locations and site sizes of distamycin and penta-N-methylpyrrolecarboxamide on three DNA restriction fragments from pBR322 plasmid. We find that methidiumpropyl-EDTA.Fe(II) footprinting and DNA affinity cleaving methods report common binding locations and site sizes for the tri- and pentapeptides bound to heterogeneous DNA. The tripeptide distamycin binds 5-base-pair sites with a preference for poly(dA).poly(dT) regions. The pentapeptide binds 6-7-base-pair sites with a preference for poly(dA).poly(dT) regions. These results are consistent with distamycin binding as an isogeometric helix to the minor groove of DNA with the four carboxamide N-H's hydrogen bonding five A + T base pairs. The data supports a model where each of the carboxamide N-H's can hydrogen bond to two bases, either O(2) of thymine or N(3) of adenine, located on adjacent base pairs on opposite strands of the helix. In most (but not all) cases the tri- and pentapeptide can adopt two orientations at each A + T rich binding site.  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