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.
The changes in polymer-solvent interactions that occur when native calf thymus DNA is dialyzed against Na2SO4 solutions of a given ionic strength and buffer concentration but of varying concentrations in methylmercuric hydroxide have been investigated with the help of solution density measurements at 25 °C and pH 6.8–7.0. From measurements executed under equilibrium dialysis conditions at the three salt levels 5 mm, 0.05 m, and 0.5 m Na2SO4 (m refers to molality) and in the presence of 5 mm cacodylic acid buffer, the density increments (???c2)μ0 for native calf thymus DNA were determined as a function of CH3HgOH concentration. (???c2)μ0 was found not to vary with organomercurial concentration, irrespective of the concentration of supporting electrolyte, until a certain CH3HgOH concentration level has been reached, viz., pM1 ? 3.5 (pM1 = ?log mCH3HgOH), beyond which (???c2)μ0 increases strongly with increasing concentration of CH3HgOH. As is shown by optical melting, (???c2)μ0 becomes a function of organomercurial concentration the moment DNA undergoes denaturation brought about by the complexing of CH3HgOH with the various N-binding sites of the base residues in the DNA double helix.Polymer-solvent interactions, expressed in terms of preferential water interactions (“net hydration”) and preferential salt interactions (“salt solvation”), were derived from the (???c2)μ0 data in combination with data obtained on the preferential interaction of CH3HgOH with denatured DNA and data on the partial specific volumes of all major solution components, gathered from density measurements on solutions with fixed concentrations of diffusible components. Evidence is presented which shows that denaturation in general decreases the net hydration while salt becomes preferentially associated with the polyelectrolyte. This process is further amplified by the interaction of CH3HgOH with denatured DNA: Methylmercurated DNA alters the redistribution of diffusible components at dialysis equilibrium to such an extent that in a formal sense large amounts of water are rejected from the immediate vicinity of the polymer. The molecular implications of these findings are explored. The results are further discussed in the light of previous findings where the methylmercury-induced denaturation of DNA had been studied with the help of buoyant density measurements in a Cs2SO4 density gradient and by velocity-sedimentation in a variety of sulfate media.  相似文献   

2.
M J Tunis  J E Hearst 《Biopolymers》1968,6(9):1325-1344
The hydration of DNA is an important factor in the stability of its secondary structure. Methods for measuring the hydration of DNA in solution and the results of various techniques are compared and discussed critically. The buoyant density of native and denatured T-7 bacteriophage DNA in potassium trifluoroacetate (KTFA) solution has been measured as a function of temperature between 5 and 50°C. The buoyant density of native DNA increased linearly with temperature, with a dependence of (2.3 ± 0.5) × 10?4 g/cc-°C. DNA which has been heat denatured and quenched at 0°C in the salt solution shows a similar dependence of buoyant density on temperature at temperatures far below the Tm, and above the Tm. However, there is an inflection region in the buoyant density versus T curve over a wide range of temperatures below the Tm. Optical density versus temperature studies showed that this is due to the. inhibition by KTFA of recovery of secondary structure on quenching. If the partial specific volume is assumed to be the same for native and denatured DNA, the loss of water of hydration on denaturation is calculated to be about 20% in KTFA at a water activity of 0.7 at 25°C. By treating the denaturation of DNA as a phase transition, an equation has immmi derived relating the destabilizing effect of trifluoroacetate to the loss of hydration on denaturation. The hydration of native DNA is abnormally high in the presence of this anion, and the loss of hydration on denaturation is greater than in CsCl. In addition, trifluoroacetate appears to decrease the ΔHof denaturation.  相似文献   

3.
The equilibrium and the stoichiometry for the reversible complexing of silver ion by DNA have been studied by potentiometric titrations, proton release pH-stat titrations, and by spectrophotometry. The complexing reactions involve primarily the purine and pyrimidine residues, not the phosphate groups. There are at least three types of binding (types I, II, and III), of which the first two have been intensively studied in this work. The sum of type I and type II binding saturates at one silver atom per two nucleotide residues. In the type I and type II reactions, zero and one proton, respectively, are displaced per silver ion bound. At pH 5.6, the reactions occur stepwise, type I being first, while at pH 8.0, they occur simultaneously. The silver ion binding curve is very sharp at pH 8, indicating a cooperative reaction. The strength of the binding increases with increasing GC content. Type I binding is more important for GC-rich DNA's than for GC-poor ones. Denatured DNA binds more strongly than does native DNA. The silver ion complexing reaction is chemically and biologically reversible. We propose that type II binding essentially involves the conversion of an \documentclass{article}\pagestyle{empty}\begin{document}$ {\rm N} - {\rm H} \cdots {\rm N} $\end{document} hydrogen bond of a complementary base pair to an N—Ag—N bond. The nature of type I binding is less clear, but it may involve a π interaction with stacked bases. The buoyant density (ρ0) of DNA in a Cs2SO4 density gradient increases when the DNA reacts with silver ion. The buoyant density change is about 0.15 g./ml. for 0.5 silver bound per nucleotide. The large buoyant density changes and the selective nature of the complexing reaction make it possible to perform good separations between native and denatured DNA or between GC-rich and GC-poor native DNA's by density gradient centrifugation.  相似文献   

4.
The effects of increasing concentartions of methylmercuric hydroxide (CH3HgOH) on the rate of sedimentation, S0, and intrinsic viscosity, [η], of T7 DNA have been studied at 20°C in 0.005, 0.05, and 0.5M Na2SO4, respectively, whereby each salt solvent conatined, in addition, 0.005M sodium borate, pH 9.18, as a buffer. Both S0 and [η] are independent of organomercurial concentration as long as DNA remains native. Denaturation, brought about by the complexing of CH3HgOH with the polymer, produces large changes in S0 as wll as [η]. The sedimentation coefficient increases strongly with increasing oragnomercurial concentration once strand separation has occured. Experimental difficulties prevented measuring of [η] in the posttransition region. The data on S0 have been used, in combination with available information on the so-called density increment (?ρ/?c2), to obtain information on the frictional properties of single-stranded and methylmercurated T7 DNA. The frictional coefficient, defined as f′2 = M2(?p/?c2)/S0ηNA, where M2 is the molecular wieght of T7 DNA, c2 is the concentration of DNA in g/ml of solution, ηr the realtive viscosity of the salt solvents, and NA is Avogadro's number, was evaluated for all three salt media as a function of organomercurial concentration. f′2 of native T7 DNA was found not to be sensitive to changes in ionic strngth; but f′2 of single-stranded and methylmercurated T7 DNA varied strongly with salt concentration. Since f′2 of single-stranded T7 DNA was barely affected by organomercurial concentration at a given ionic strength, it is concluded that the dramatic variations of S0 with pM (pM ≡ -log[CH3HgOH]) observed in the posttransition zone reflect only changes in the thermodynamic interactions (“preferential interactions”) existing between DNA and the vatious other solution components, but not changes in the coil dimensions of the polymer.  相似文献   

5.
HeLa S3 cells were exposed to varied concentrations of methylmercury over varied periods of time and its binding by the cells was studied using 203Hg-labeled methylmercuric chloride as radioactive marker. Also studied was the effect of cell-bound methylmercury on DNA replication and protein synthesis and on the growth rate of the cells. The results show that methylmercury binding is a rapid process, with much of the organomercurial bound within the the first 60 min of incubation, and that considerable quantities of organic mercury become affixed to the cells. The amounts of bound methylmercury, [CH3Hg(II)]bound, given in mol/cell, range from 2 × 10?16 (at 1 h of incubation and at 1 μM CH3Hg(II) in the medium) to almost 4 × 10?14 (at 24 h of incubation and at 100 μM CH3Hg(II) in the medium). A [CH3Hg(II)]bound value of about 30 × 10?16 mol/cell appears to be the threshold below which cells display a normal growth pattern and below which metabolic events such as DNA replication or protein synthesis are affected only to a minor degree but above which major changes in cell metabolism and cell growth take place. Methylmercury binding by the cells is tight so that only 20% of the bound material is released from the cells over a 3-h incubation period when the cells are placed into fresh, methylmercury-free growth medium. Analysis of the binding data in terms of binding to identical and completely independent sites yields an association constant K of 7.92 × 104 l/mol and for the maximum concentration of cellular binding sites the value 2.40 × 10?14 mol/cell or 1.45 × 1010 sites/cell. Evidence is presented which shows that cellular sulfhydryl groups do not suffice to provide all the sites taken up by methylmercury and that binding, in all likelihood, involves basic nitrogen, too. The levels of cell-bound methylmercury are such that binding to HeLa DNA and HeLa chromatin, for instance, can readily take place. Methylmercury binding data obtained by using the technique of particle-induced X-ray emission (PIXE) are in good agreement with the data obtained via isotope dilution.  相似文献   

6.
A simple and rapid procedure for recovering the denaturing effect of methylmercuric hydroxide in agarose gel electrophoresis is described. The procedure consisted of the treatment of the commercial methylmercuric hydroxide solutions with Amberlite, a mixture of anion- and cation-exchange resins. This treatment greatly improved the resolution of RNA species when fractionated by electrophoresis through agarose-CH3HgOH slab gel.  相似文献   

7.
A previous study in our laboratory of the reaction of formaldehyde with super-helical DNAs (φX replicative form and PM2) has led to a model for superhelical DNA in which there is a region or regions of altered secondary structure containing unpaired bases. Similar experiments using the nicked circular DNA gave no evidence of interruptions of base pairing. In this study we present additional data, which support the above model as well as extending our analysis of the secondary structure of superhelical DNA and the dynamics of the early denaturation process. In a series of experiments involving the binding of methyl-mercury as a chemical probe of unpaired bases, we obtained the following results. (1) Initially, both s020w and the buoyant density of the superhelical form of phage PM2 DNA increased as a function of methylmercuric hydroxide concentration, whereas the nicked form did not. (2) This initial binding is accompanied by an increase in superhelical content τ from ?41 to ?46 turns. (3) The binding analysis allows us to estimate that 3.7% of the bases contain methylmercury in this phase of the transition. This is in excellent agreement with the extent of formylation. (4) Such a preformylated molecule shows a shift in the transition to lower mercurial concentrations. These results are interpreted as follows. The initial increase in ?τ excludes the possibility that binding occurs to normal base-paired structures, since this would produce a coupled unwinding of duplex and superhelical turns. The additive effects of formylation and methylmercury binding support the concept that both chemical probes attack the same sites and induce similar structural changes. Thus the evidence clearly supports the view that superhelical DNA contains localized region(s) of interrupted base pairing. Recent studies from other laboratories using single strand-specific endonucleases are in complete agreement with this model.  相似文献   

8.
Dilatometric measurements were made to determine the change in apparent specific volume φ of DNA resulting from thermal denaturation in neutral solution, φ increased continuously with temperature in the range 10–85°C. No deviations from a monotonically rising curve were observed in the φ versus temperature profile in the region of the melting temperature. The results are interpreted in terms of a partial loss of the preferentially bound DNA hydration shell. The nature of the well known buoyant density difference between native and denatured DNA was investigated by evaluating the densities in a series of cesium salt gradients at constant temperature. Extrapolation of the results to zero water activity indicates that the partial specific volumes of anhydrous native and denatured DNA are equal. The density difference at nonzero water activities is attributed to decreased hydration in the denatured state. The absence of a related change in φ accompanying the denaturation in the dilatometric experiments suggests that the probable volume change associated with loss of bound water during denaturation is accompanied by other compensatory volume effects. The possible nature of these volume effects is discussed.  相似文献   

9.
10.
Aqueous RbTCA is generally suitable as a buoyant solvent for both native and denatured DNA at neutral pH and room temperature. Native PM-2 DNA II, for example, is buoyant at 3.29 M salt, 25 degrees C; whereas the denatured strands band together at 4.52 M. Two properties of the solvent make this system uniquely useful for separations based upon the extent of secondary structure. First, the melting transition temperature for chemically unaltered DNA is depressed to room temperature or below. Second, the buoyant density increase accompanying denaturation is extraordinarily large, 174 mg/ml for PM-2 DNA II. This value is three times that found in aqueous NaI and ten times that for CsCl. The properties of the RbTCA buoyant solvent presented here include the compositional and buoyant density gradients and the buoyant density dependence upon base composition. The DNA remains chemically unaltered after exposure to RbTCA as shown by the absence of strand scissions for closed circular DNA and by the unimpaired biological activity in transformation assays. Intact virion DNA may be isolated by direct banding of whole virions in RbTCA gradients without prior phenol extraction. Strongly complexed or covalently bound proteins may be detected by their association with the buoyant polymer in the denaturing density gradient.  相似文献   

11.
Using a combination of As and Se K-edge and Hg LIII-edge X-ray absorption spectroscopy, 77Se nuclear magnetic resonance spectroscopy, electrospray ionization mass spectrometry and molecular modeling, we have structurally characterized the novel species methylmercury(II) seleno bis(S-glutathionyl) arsenic(III). This species is formed in aqueous solution from CH3HgOH and the seleno bis(S-glutathionyl) arsinium ion and constitutes an important first step towards characterizing the observed toxicologically relevant interaction between arsenite, selenite and methylmercury which has been previously reported in mammals.  相似文献   

12.
Northern blotting using glyoxal to denature a highly folded mRNA, such as goat αs1-Casein E, can lead to the detection of multiple incompletely denatured forms. Formaldehyde appears to be the most suitable agent for Northern blotting due to its effective denaturing capacity and lower toxicity than methylmercuric hydroxide.  相似文献   

13.
C-reactive protein (CRP) mRNA was assayed by cell-free translation of poly(A)-containing liver RNA isolated both from rabbits stimulated to undergo the acute-phase response and from unstimulated control rabbits. No CRP-related translation products were identified until the denaturant methylmercury hydroxide (CH3HgOH) was added to the RNA before cell-free translation. In the presence of the denaturant, a 24000-Da translation product was synthesized which was immunochemically identifiable as the CRP primary translation product. It is likely that rabbit CRP mRNA can form a stable intramolecular duplex which interferes with its translatability in vitro. The 24000-Da CH3HgOH-facilitated cell-free translation product was not detected in poly(A)-containing liver RNA from unstimulated animals, indicating that the concentration of translatable CRP mRNA was dramatically induced during the acute-phase response. On the basis of absorption experiments, the 24000-Da CRP primary translation product was immunochemically more closely related to denatured CRP than to native CRP.  相似文献   

14.
H.C. Birnboim  R. Sederoff 《Cell》1975,5(2):173-181
Very long runs of pyrimidine nucleotides (polypyrimidines), previously detected in DNA from Drosophila melanogaster, have now been localized to a “cryptic” satellite. These polypyrimidines have an average length of 750 nucleotides and account for about 3% of the thymine residues in total DNA. The buoyant density of the DNA component which contains the polypyrimidines was detected by centrifuging native DNA to equilibrium in a CsCI gradient, and then assaying each fraction for its content of polypyrimidines. A peak was detected at a density of about 1.707 gm/cm3, distinctly heavier than the main band of DNA (1.702 gm/cm3). The buoyant density of polypyrimidine-containing molecules was little affected by differences in the molecular weight of the starting DNA in the range 105-107 daltons (single-stranded). Thus polypyrimidines (and their complementary polypurines) appear to form all or part of a “cryptic” satellite.Polypyrimidines have been isolated and characterized with respect to composition and buoyant density. Direct nucleoside analysis of unlabeled material indicated 34.5% deoxycytidine, 65.5% thymidine. Their banding position in neutral and alkaline CsCI gradients was consistent with a single-stranded DNA polymer of this composition.  相似文献   

15.
The arrangement of the reiterated DNA sequences complementary to transfer RNA has been studied in Xenopus laevis. Prehybridization of denatured DNA with an excess of unfractionated tRNA results in a small but well-defined increase in the buoyant density of fragments which contain sequences homologous to tRNA. The density increase is smaller than that found for 5 S DNA, but is the same or nearly so for all tRNA coding sequences examined. These results indicate that the majority of tRNA genes are clustered together with spacer DNA, the average size of which is estimated to be approximately 0.5 × 106 daltons (native) DNA.In high molecular weight native DNA preparations, the sequences homologous to unfractionated tRNA, tRNAVal, tRNA1Met and tRNA2Met band in CsCl at 1.707, 1.702, 1.708 and 1.711 g cm?3, respectively. The mean buoyant densities are constant at all molecular weights examined but they do not correspond to the base compositions of the complementary tRNA species. These results indicate that isocoding genes are linked to spacer DNA in separate and extensive gene clusters, and that the different clusters contain different spacer DNA sequences. These clusters form well-defined cryptic DNA satellites which are potentially separable from each other as well as from other chromosomal DNA.  相似文献   

16.
The cytoplasmic DNA-binding proteins of Physarum polycephalum were recovered by chromatography of cytosol extracts on sequential columns of native and denatured calf thymus DNA-cellulose. 5.4% of the total cytosol protein was bound to native DNA-cellulose, while 4.4% was bound to denatured DNA-cellulose. Stepwise salt gradient elution of the columns separated the DNA-binding proteins into 9 fractions which were analysed by acrylamide gel electrophoresis in the presence of sodium dodecyl sulfate. Several hundred discrete polypeptide bands were identified, with many more high molecular weight polypeptides (greater than 100 000 D) binding to native than to denatured DNA. Continuous in vivo labelling of microplasmodia in KH2[32P]O4 and [3H]leucine was used to determine which of the DNA-binding proteins were phosphorylated, and to approximate their phosphorus content. About 30–40 phosphoproteins were resolved among the DNA-binding proteins. Most phosphoproteins contained less than 3 phosphates per polypeptide, but a small number of low molecular weight phosphoproteins (less than 50 000 D) contained from 5 to 10 phosphates per polypeptide. The majority of high molecular weight DNA-binding phosphoproteins bound to native DNA and were eluted with 0.25 M NaCl. As a group, the DNA-binding proteins were enriched in protein-bound phosphorus when compared with the cytosol proteins which did not bind to DNA. The phosphorus content of the cytoplasmic DNA-binding proteins was similar to that of the acidic nuclear proteins.  相似文献   

17.
Four genomic DNAs of differing GC content (Micrococcus luteus, 72% GC; Escherichia coli, 50% GC; calf thymus, 42% GC; Clostridium perfringens, 27% GC) have been employed as targets of interaction by the cationic polyamines spermidine {[H3N(CH2)3NH2(CH2)4NH3]3+} and spermine {[(CH2)4(NH2(CH2)3NH3)2]4+}. In solutions containing 60 mM DNA phosphate (~20 mg DNA/ml) and either 1, 5 or 60 mM polyamine, only Raman bands associated with the phosphates exhibit large spectral changes, demonstrating that B-DNA phosphates are the primary targets of interaction. Phosphate perturbations, which are independent of base composition, are consistent with a model of non-specific cation binding in which delocalized polyamines diffuse along DNA while confined by the strong electrostatic potential gradient perpendicular to the helix axis. This finding provides experimental support for models in which polyamine-induced DNA condensation is driven by non-specific electrostatic binding. The Raman spectra also demonstrate that major groove sites (guanine N7 and thymine C5H3) are less affected than phosphates by polyamine–DNA interactions. Modest dependence of polyamine binding on genome base composition suggests that sequence context plays only a secondary role in recognition. Importantly, the results demonstrate that polyamine binding has a negligible effect on the native B-form secondary structure. The capability of spermidine or spermine to bind and condense genomic B-DNA without disrupting the native structure must be taken into account when considering DNA organization within bacterial nucleoids or cell nuclei.  相似文献   

18.
The effects of temperature and ethidium bromide on the banding of heat-denatured DNA was studied during equilibrium centrifugation in density gradients of NaI. Centrifugation at 10 degrees C prevents the partial renaturation of Escherichia coli DNA and Clostridium perfringens DNA that occurs at 20 degrees C. A centrifugation temperature of --5 degrees C is required to prevent renaturation of T7 phage DNA. Ethidium bromide decreases renaturation of Escherichia coli DNA during centrifugation at 20 degrees C and causes a small shift in the buoyant density of both denatured and native DNA. Equilibrium centrifugation at lower temperatures prevents DNA renaturation and permits increased utilization of the large buoyant density difference between native and heat-denatured DNA in gradients of NaI.  相似文献   

19.
Nuclei from hamster embryo fibroblasts treated with radioactive benzo(a)pyrene were lysed in 6 m guanidine, and nuclear macromolecules were separated by isopycnic centrifugation in Cs2SO4. Control experiments showed that cross-contamination of the RNA, DNA, and protein fractions was less than 2% of the total recovery of each macromolecular class. When compared to previous techniques utilizing phenol extraction, similar specific activities of bound hydrocarbon (pmol benzo[a]pyrene/mg protein or nucleic acid) were obtained. However, overall recoveries of macromolecular components were higher with the present method. In addition, recovery of undegraded histones in the density gradient preparation of nuclear protein was demonstrated by sodium dodecyl sulfate-polyacrylamide gel electrophoresis, and recovery of native DNA was demonstrated by thermal denaturation studies. Although developed specifically for work with carcinogenic hydrocarbons, the Cs2SO4 technique should be generally useful in cases where it is necessary to prepare all three classes of macromolecules from one batch of nuclei.  相似文献   

20.
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   

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