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 31 毫秒
1.
Monomolecular condensation of lambda-DNA induced by cobalt hexamine   总被引:9,自引:0,他引:9  
J Widom  R L Baldwin 《Biopolymers》1983,22(6):1595-1620
Measurements of static and dynamic light scattering have been used to distinguish between monomolecular DNA condensation and aggregation of condensed molecules. In low salt, using Co3+(NH3)6 as the condensing agent, and at λ-DNA concentrations below 0.2 μg/mL, the transition curves for monomolecular condensation and aggregation are well separated for times of 16 h. In these conditions, the intensity of scattered light (90°) and also the diffusion coefficient of the condensed DNA show reasonable values for monomolecular condensation that are independent of DNA concentration and also of Na+ Co3+(NH3)6 concentrations for which monomolecular condensation is complete. At higher Co3+(NH3)6 concentrations, which produce aggregation (as judged by the intensity of scattered light), the diffusion coefficient decreases sharply. The transition curve for monomolecular condensation is independent of DNA concentration but shows a hysteresis loop. The kinetics of condensation are slow in the forward direction and fast in the reverse direction, indicating that the actual transition curve is measured closely by reversal experiments. Aggregation is blocked kinetically in both the forward and reverse directions when Co3+(NH3)6 is the condensing agent at low Na+ concentrations. When spermine or spermidine is the condensing agent and observations are made at 16 h, it is not possible to separate the transition curves for monomolecular condensation and for aggregation in conditions that are successful with Co3+(NH3)6. Some interesting properties of monomolecular condensation are noted. (1) The transition is not a two-state reaction, as judged by measurements of the diffusion coefficient through the transition zone. (2) The transition for monomolecular condensation is diffuse. (3) The dimensions of the monomolecular condensates have been calculated from the translational diffusion coefficient for an assumed toroidal shape by the formula derived by Allison and coworkers [(1981) Biopolymers 20 , 469–488]. These dimensions are in reasonable agreement with ones deduced from electron microscopy by Chattoraj and coworkers [(1978) J. Mol. Biol. 121 , 327–337]. (4) The phase diagram relating the Na+ to the Co3+(NH3)6 concentrations needed for condensation has a slope of 0.6 in a log–log plot. According to numerical solutions of Manning's theory for the atmospheric binding of competing cations to DNA, this means that condensation occurs at a late stage in the replacement of Na+ by Co3+(NH3)6 around the DNA. The fraction of DNA phosphate charge neutralized at condensation is computed to be in the neighborhood of 0.9, as found by Wilson and Bloomfield [(1979) Biochemistry 18 , 2192–2196], but to vary with the Na+ concentration.  相似文献   

2.
3.
Abstract

The condensation of chicken erythrocyte (CE) and calf thymus (CT) chromatins upon addition of di- and multivalent cations has been studied using turbidityJulprecipitation and electric dichroism measurements. For all the cations investigated (Mg2+, Tb3+, Co(NH3)6 3+, spermidine Spd2+ and spermine Sp4+) condensation of CE chromatin occurred before the onset of aggregation, while aggregation of CT chromatin started before condensation with all cations except Mg2+ and Tb3+. Precipitation of CE chromatin required lower di- and multivalent cations concentrations than CT chromatin. The electric dichroism data for both chromatins, at low ionic strength in the absence of di- or multivalent cations, indicated that the nucleoprotein molecules were not totally decondensed but that a “precondensed” state was already present. A positive electric dichroism was observed for the most condensed chromatin fibers, in agreement with the “cross-linker” models. Tb3+ led to less compact condensed particles as judged from the electric dichroism observations, but electron microscopy revealed that “30 nm fibers” were formed. Very little aggregation was produced by Tb3+. On the contrary, spermine produced very large networks of condensed molecules, but large spheroidal particles were also observed. The condensation of CE chromatin happened without changes of solution conductivity upon cation salt addition, regardless of the condensing cation, indicating a cooperative uptake of the ions during this process.  相似文献   

4.
D Porschke 《Biopolymers》1985,24(10):1981-1993
Electric-field pulses of e.g. 20 kV/cm and 100 μs induce a strong decrease in the scattered light intensity of DNA condensed by spermine. Analysis of this effect demonstrates that the decrease of the scattered light intensity results from decondensation of DNA. The decondensation reaction requires an electric-field strength exceeding a threshold value. Complete decondensation can be achieved at field strength that are only slightly higher than the threshold value. The decondensation process is strongly accelerated at high electric-field strengths. At 30 kV/cm, the decondensation time constant is ~8 μs, corresponding to an acceleation factor of 105 relative to the field-free decondensation reaction. The dependence of the time constants on the electric-field strength suggests that the field-induced decondensation is due to a dissociation field effect. The condensation process observed after electric-field pulses at low concentrations of DNA and spermine shows a characteristic induction period, which strongly depends on the spermine concentration. This induction period reflects the time required for the binding of spermine to DNA, until the degree of binding is sufficiently high for the condensation reaction. The fast dissociation of condensed DNA by electric-field pulses together with a relatively long lifetime of the free DNA results in a reaction cycle resembling a hysteresis loop.  相似文献   

5.
We have measured forces generated by multivalent cation-induced DNA condensation using single-molecule magnetic tweezers. In the presence of cobalt hexammine, spermidine, or spermine, stretched DNA exhibits an abrupt configurational change from extended to condensed. This occurs at a well-defined condensation force that is nearly equal to the condensation free energy per unit length. The multivalent cation concentration dependence for this condensation force gives the apparent number of multivalent cations that bind DNA upon condensation. The measurements show that the lower critical concentration for cobalt hexammine as compared to spermidine is due to a difference in ion binding, not a difference in the electrostatic energy of the condensed state as previously thought. We also show that the resolubilization of condensed DNA can be described using a traditional Manning–Oosawa cation adsorption model, provided that cation–anion pairing at high electrolyte concentrations is taken into account. Neither overcharging nor significant alterations in the condensed state are required to describe the resolubilization of condensed DNA. The same model also describes the spermidine3+/Na+ phase diagram measured previously.  相似文献   

6.
The interaction of ethidium bromide, a DNA intercalating drug, and bis( methidium )spermine, a DNA bis-intercalating compound, with the left-handed Z form of poly(dG-dC) has been studied in 4.4 M NaCl. Spectrophotometric analysis using absorption, fluorescence and circular dichroism indicates that the complex formed between ethidium and Z DNA resembles very closely that formed with B DNA. This suggests that ethidium binds to Z DNA by intercalation. 31P NMR spectra are presented showing both the conversion of the Z form to the B form with increasing amounts of drug and the typical Z form spectrum at low binding densities. Data are also presented which show that the bifunctional intercalator bis( methidium )spermine binds to Z DNA in a manner similar to its binding to B DNA, i.e., by bis-intercalation. These results are important for our understanding the behavior of Z DNA and its biological significance.  相似文献   

7.
The chemically inert trivalent ion cobalt(III) hexaammine, Co3+(NH3)6, was found to exert polyamine-like effects in enhancing certain cyclic nucleotide-independent protein kinase reactions catalyzed by nuclear enzyme preparations from rat ventral prostate or liver. At 1 mM, Co3+(NH3)6 stimulated chromatin- and also non-histone-protein-associated kinase activities with partially dephosphorylated phosvitin as substrate by 38% and 72% respectively, whereas chromatin-associated kinase-catalyzed phosphorylation of lysine-rich histones was not affected under the same conditions. 32P incorporation (from γ-32P-ATP) into endogenous protein substrates of chromatin or non-histone protein fractions catalyzed by their erdogencus kinase activity was increased by 47% and 153%, respectively. These effects of Co3+(NH3)6 were similar to those produced by 1mM spermine. Autoradiographic analysis of endogenous 32P-labelled nonhistone proteins revealed similar enhancements of the phosphorylation of several of the same proteins, induced by 1mM spermine or 1 mM Co3+(NH3)6 or 2mM spermidine. The stimulatory actions of polyamines or Co3+(NH3)6 were not mimicked by raising the ionic strength by addition of comparable concentrations of NaCl. The effects of 1 mM spermine and of 1 mM Co3+(NH3)6 tested separately were not additive. Phosphorylation of lysine-rich histones by beef heart cyclic AMP-dependent protein kinase was not affected by polyamines or Co3+(NH3)6 Various findings hint that the enhancement of cyclic nucleotide-independent kinase-catalyzed phosphorylation of certain protein substrates by spermidine, spermine and Co3+(NH3)6 is primarily due to interaction of these cations with appropriate protein substrates affecting their conformational status. Further, these effects of polyamines may be a reflection of their cationic charge properties rather than being dependent on any particular conformations assumed by the polyamines.  相似文献   

8.
Chromomycin (Chro) forms a 2:1 drug/metal complex through the chelation with Fe(II), Co(II), or Cu(II) ion. The effects of spermine on the interaction of Fe(II), Co(II), and Cu(II) complexes of dimeric Chro with DNA were studied. Circular dichroism (CD) measurements revealed that spermine strongly competed for the Fe(II) and Cu(II) cations in dimeric Chro-DNA complexes, and disrupted the structures of these complexes. However, the DNA-CoII(Chro)2 complex showed extreme resistance to spermine-mediated competition for the Co(II) cation. According to surface plasmon resonance (SPR) experiments, a 6 mM concentration of spermine completely abolished the DNA-binding activity of FeII(Chro)2 and CuII(Chro)2 and interfered with the associative binding of CoII(Chro)2 complexes to DNA duplexes, but only slightly affected dissociation. In DNA integrity assays, lower concentrations of spermine (1 and 2 mM) promoted DNA strand cleavage by CuII(Chro)2, whereas various concentrations of spermine protected plasmid DNA from damage caused by either CoII(Chro)2 or FeII(Chro)2. Additionally, DNA condensation was observed in the reactions of DNA, spermine, and FeII(Chro)2. Despite the fact that CuII(Chro)2 and FeII(Chro)2 demonstrated lower DNA-binding activity than CoII(Chro)2 in the absence of spermine, while CuII(Chro)2 and FeII(Chro)2 exhibited greater cytoxicity against HepG2 cells than CoII(Chro)2, possibly due to competition of spermine for Fe(II) or Cu(II) in the dimeric Chro complex in the nucleus of the cancer cells. Our results should have significant relevance to future developments in metalloantibiotics for cancer therapy.  相似文献   

9.
S M Cheng  S C Mohr 《Biopolymers》1975,14(3):663-674
Circular dichroism spectroscopy has been used to investigate the influence of DNA molecular size, base composition, and the presence of intercalating agents upon the Ψ transition of DNA brought about by high concentrations of poly(ethylene oxide) and salt (Lerman (1971) Proc. Natl. Acad. Sci. (U.S.) 68 , 1886–1890). A molecular weight of 0.15–3.0 × 106 daltons yields maximum formation of Ψ-DNA. Both the amplitude of the large negative CD band at 265 nm—a chief characteristic of the Ψ state—and the thermal stability of Ψ-DNA increase linearly with increasing mole fraction of guanine plus cytosine in the DNA sample. Either ethidium or proflavine, at concentrations where approximately one dye is bound per 5–10 nucleotide residues, can prevent the transition completely. Striking similarities between the Ψ-DNA produced by poly(ethylene oxide) + salt and the complexes formed between DNA and lysine-rich histone f1 suggest the presence of similar nucleic acid–nucleic acid interactions in both types of condensed phase.  相似文献   

10.
DNA molecules condense into compact structures in the presence of a critical concentration of multivalent cations. To probe the contribution ofelectrostatic forces to condensation, we used mixtures of water with methanol (MeOH), ethanol (EtOH), and isopropanol (iPrOH) to vary the dielectric constant ? from 80 to 50. The condensation of pUC18 plasmids by hexaammine cobalt (III), Co(NH3), was monitored by total intensity and dynamic light scattering, electron microscopy, andCD. The total scattering intensity increased as ? went from 80 to 70, and then decreased as ? decreased further. Ultraviolet spectrophotometry confirmed that the loss of intensity at low ? was not due to the particles' settling out of solution. The rate as well as the extent of condensation increased as? was lowered from 80 to 70, and also depended on the species of alcohol (MeOH < EtOH < iPrOH). The hydrodynamic radii RH of the particles, however, remained roughly the same at 300–350 A and was independent of the species of alcohol. RH increased below ? = 70. The critical concentration of Co(NH3) required to induce DNA condensation decreased from 21 μM to about 16 μM as the dielectric constant decreased from 80 to 70, and decreased moderately with the nonpolarity of the alcohol. The fraction of DNA charge neutralized at the onset of DNA condensation was calculated by a modification of Manning's two-variable counterion condensation theory to be 0.90 ± 0.01, independent of ?. By electron microscopy we observed that the condensed particles changed from about 93% toroids at ? = 80 to 89% rods at ? = 70 and 98% rods at ? = 65. At epsi; lower than 65, DNA collapsed into a network of multistranded fibers. The morphology of condensed DNA particles, whether toroids, rods, or fibers, was independent of the alcohol species. CD spectra in ethanol–water mixtures indicated that both closed circular and linearized plasmids were in the B conformation when condensed with Co(NH3)3+6 at ?≥ 70, although the closed circular molecules exhibited a weak Ψ-DNA spectrum. A transition from the B to A formtook place between ? = 70 and 60, well above the normal dielectric constant of ? = 40 for this transition, indicating that ethanol and Co(NH3) synergistically promote the B–A transition. We interpret these results to mean that alcohols have both electrostatic and structural effects on DNA, leading to three regimes of condensation. At the lowest alcohol concentrations the B conformation is stableand condensation is relatively slow, allowing time for the packing adjustments necessary to form toroids. At intermediate alcohol concentrations condensation is faster, and the combined effects of solvent and Co(NH3) locally destabilize the double helix, permitting DNA foldbacks that lead to rodlike condensates. Rods become shorter as wellas more numerous as ? decreases from 80 to 65–60, indicatingincreasing destabilization as alcohol increases. At the lowest dielectric constants, alcohol and Co(NH3) produce A-DNA, which strongly self-adheres and rapidly aggregates intofibrous networks, not allowing time for more compact condensates to form. © 1995 John Wiley & Sons, Inc.  相似文献   

11.
We synthesized a pentamine (3-3-3-3) and two hexamine (3-3-3-3-3 and 3-4-3-4-3) analogues of the natural polyamine, spermine (3-4-3) and studied their effectiveness in condensing pGL3 plasmid DNA, using light scattering and atomic force microscopic (AFM) techniques. The midpoint concentration of the polyamines on pGL3 condensation (EC50) was 11.3, 10.6, 1.5, 0.49 and 0.52 µM, respectively, for 3-4-3, norspermine (3-3-3), 3-3-3-3, 3-3-3-3-3 and 3-4-3-4-3 in 10 mM Na cacodylate buffer. Dynamic laser light scattering study showed a decrease in hydrodynamic radii of plasmid DNA particles as the number of positive charges on the polyamines increased. AFM data showed the presence of toroids with outer diameter of 117–191 nm for different polyamines, and a mean height of 2.61 ± 0.77 nm. AFM results also revealed the presence of intermediate structures, including those showing circumferential winding of DNA to toroids. The dependence of the EC50 on Na+ concentration suggests different modes of binding of spermine and its higher valent analogues with DNA. Our results show a 20-fold increase in the efficacy of hexamines for DNA condensation compared to spermine, and provide new insights into the mechanism(s) of DNA nanoparticle formation. These studies might help to develop novel nonviral gene delivery vehicles.  相似文献   

12.
Interactions of certain intercalating cationic ligands with nucleic acids result in the formation of products that undergo condensation and agglomeration; this transition in solution can be monitored by light-scatter measurements. In the present study, using such intercalators as the antitumor drug mitoxantrone or fluorochromes acridine orange and quinacrine, we induced condensation of DNA in situ in Chinese hamster chromosomes. The in situ products scattered light and could be detected by darkfield- or phase-contrast microscopy. In the darkfield the complexes had a characteristic granular appearance and often generated a banding pattern on the chromosomes. In contrast, condensation of DNA in situ by the nonintercalating polyvalent cations (Co3+, spermine4+), while enhancing the chromosome's image contrast, did not produce the granular products or the banding. The condensation of free DNA, single or double stranded, natural or synthetic, the latter of various base composition and configuration, was also measured in solution. The condensation in solution and in situ was observed at similar concentrations of the respective ligands. The intercalating dye ethidium bromide, which did not condense DNA in solutions of moderate and high ionic strength, also did not generate the granular products or banding on chromosomes. The data also show that both base composition and configuration are important factors in determining the sensitivity of DNA to condensation by particular intercalating ligands. The studies suggest that the phenomenon of DNA condensation by intercalating dyes, which shows a high degree of specificity with respect to primary and secondary structures of DNA, may be associated with mechanisms of chromosome banding induced by the intercalating thiazine dyes in Giemsa staining or by quinacrine. Observation of chromosome banding based on light-scatter detection in darkfield microscopy allows the study of interactions between DNA and the ligands that neither fluoresce nor generate colored products. This principle of chromosome "counter-staining" can be explored by flow cytometry.  相似文献   

13.
The condensation of chicken erythrocyte (CE) and calf thymus (CT) chromatins upon addition of di- and multivalent cations has been studied using turbidity, precipitation and electric dichroism measurements. For all the cations investigated (Mg2+, Tb3+, Co(NH3)6(3+), spermidine Spd2+ and spermine Sp4+) condensation of CE chromatin occurred before the onset of aggregation, while aggregation of CT chromatin started before condensation with all cations except Mg2+ and Tb3+. Precipitation of CE chromatin required lower di- and multivalent cations concentrations than CT chromatin. The electric dichroism data for both chromatins, at low ionic strength in the absence of di- or multivalent cations, indicated that the nucleoprotein molecules were not totally decondensed but that a "precondensed" state was already present. A positive electric dichroism was observed for the most condensed chromatin fibers, in agreement with the "cross-linker" models. Tb3+ led to less compact condensed particles as judged from the electric dichroism observations, but electron microscopy revealed that "30 nm fibers" were formed. Very little aggregation was produced by Tb3+. On the contrary, spermine produced very large networks of condensed molecules, but large spheroidal particles were also observed. The condensation of CE chromatin happened without changes of solution conductivity upon cation salt addition, regardless of the condensing cation, indicating a cooperative uptake of the ions during this process.  相似文献   

14.
The object of this study was to determine the kinetics of chromosome decondensation during the G1 period of the HeLa cell cycle. HeLa cells synchronized in the G1 period following the reversal of mitotic block were fused with Colcemid-arrested mitotic HeLa cells at 1.5, 3, 5, and 7 h after the reversal of N2O block. The resulting prematurely condensed chromosomes (PCC) were classified into six categories depending on the degree of their condensation. The frequency of occurrence of each category was plotted as a function of time after mitosis. The results of this study indicate that the process of chromosome decondensation, initiated during the telophase of mitosis continues throughout the G1 period without any interruption, thus the chromatin reaches an ultimate state of decondensation by the end of G1 period, when DNA synthesis is initiated.  相似文献   

15.
The substitution-inert polynuclear platinum(II) complex (PPC) series, [{trans-Pt(NH3)2(NH2(CH2)nNH3)}2-μ-(trans-Pt(NH3)2(NH2(CH2)nNH2)2}](NO3)8, where n = 5 (AH78P), 6 (AH78 TriplatinNC) and 7 (AH78H), are potent non-covalent DNA binding agents where nucleic acid recognition is achieved through use of the ‘phosphate clamp'' where the square-planar tetra-am(m)ine Pt(II) coordination units all form bidentate N–O–N complexes through hydrogen bonding with phosphate oxygens. The modular nature of PPC–DNA interactions results in high affinity for calf thymus DNA (Kapp ∼5 × 107 M−1). The phosphate clamp–DNA interactions result in condensation of superhelical and B-DNA, displacement of intercalated ethidium bromide and facilitate cooperative binding of Hoechst 33258 at the minor groove. The effect of linker chain length on DNA conformational changes was examined and the pentane-bridged complex, AH78P, was optimal for condensing DNA with results in the nanomolar region. Analysis of binding affinity and conformational changes for sequence-specific oligonucleotides by ITC, dialysis, ICP-MS, CD and 2D-1H NMR experiments indicate that two limiting modes of phosphate clamp binding can be distinguished through their conformational changes and strongly suggest that DNA condensation is driven by minor-groove spanning. Triplatin-DNA binding prevents endonuclease activity by type II restriction enzymes BamHI, EcoRI and SalI, and inhibition was confirmed through the development of an on-chip microfluidic protocol.  相似文献   

16.
Physical origin of DNA condensation by multivalent cations remains unsettled. Here, we report quantitative studies of how one DNA-condensing ion (Cobalt3+ Hexammine, or Co3+Hex) and one nonDNA-condensing ion (Mg2+) compete within the interstitial space in spontaneously condensed DNA arrays. As the ion concentrations in the bath solution are systematically varied, the ion contents and DNA-DNA spacings of the DNA arrays are determined by atomic emission spectroscopy and x-ray diffraction, respectively. To gain quantitative insights, we first compare the experimentally determined ion contents with predictions from exact numerical calculations based on nonlinear Poisson-Boltzmann equations. Such calculations are shown to significantly underestimate the number of Co3+Hex ions, consistent with the deficiencies of nonlinear Poisson-Boltzmann approaches in describing multivalent cations. Upon increasing the concentration of Mg2+, the Co3+Hex-condensed DNA array expands and eventually redissolves as a result of ion competition weakening DNA-DNA attraction. Although the DNA-DNA spacing depends on both Mg2+ and Co3+Hex concentrations in the bath solution, it is observed that the spacing is largely determined by a single parameter of the DNA array, the fraction of DNA charges neutralized by Co3+Hex. It is also observed that only ∼20% DNA charge neutralization by Co3+Hex is necessary for spontaneous DNA condensation. We then show that the bath ion conditions can be reduced to one variable with a simplistic ion binding model, which is able to describe the variations of both ion contents and DNA-DNA spacings reasonably well. Finally, we discuss the implications on the nature of interstitial ions and cation-mediated DNA-DNA interactions.  相似文献   

17.
Fertilization requires decondensation of promatine-condensed sperm chromatin, a dynamic process serving as an attractive system for the study of chromatin reprogramming. Nucleoplasmin is a key factor in regulating nucleosome assembly as a chaperone during fertilization process. However, knowledge on nucleoplasmin in chromatin formation remains elusive. Herein, magnetic tweezers (MT) and a chromatin assembly system were used to study the nucleoplasmin-mediated DNA decondensation/condensation at the single-molecular level in vitro. We found that protamine induces DNA condensation in a stepwise manner. Once DNA was condensed, nucleoplasmin, polyglutamic acid, and RNA could remove protamine from the DNA at different rates. The affinity binding of the different polyanions with protamine suggests chaperone-mediated chromatin decondensation activity occurs through protein–protein interactions. After decondensation, both RNA and polyglutamic acid prevented the transfer of histones onto the naked DNA. In contrast, nucleoplasmin is able to assist the histone transfer process, even though it carries the same negative charge as RNA and polyglutamic acid. These observations imply that the chaperone effects of nucleoplasmin during the decondensation/condensation process may be driven by specific spatial configuration of its acidic pentamer structure, rather than by electrostatic interaction. Our findings offer a novel molecular understanding of nucleoplasmin in sperm chromatin decondensation and subsequent developmental chromatin reprogramming at individual molecular level.  相似文献   

18.
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   

19.
Several multi-component DNA intercalating small molecules have been designed around ruthenium-based intercalating monomers to optimize DNA binding properties for therapeutic use. Here we probe the DNA binding ligand [μ-C4(cpdppz)2(phen)4Ru2]4+, which consists of two Ru(phen)2dppz2+ moieties joined by a flexible linker. To quantify ligand binding, double-stranded DNA is stretched with optical tweezers and exposed to ligand under constant applied force. In contrast to other bis-intercalators, we find that ligand association is described by a two-step process, which consists of fast bimolecular intercalation of the first dppz moiety followed by ∼10-fold slower intercalation of the second dppz moiety. The second step is rate-limited by the requirement for a DNA-ligand conformational change that allows the flexible linker to pass through the DNA duplex. Based on our measured force-dependent binding rates and ligand-induced DNA elongation measurements, we are able to map out the energy landscape and structural dynamics for both ligand binding steps. In addition, we find that at zero force the overall binding process involves fast association (∼10 s), slow dissociation (∼300 s), and very high affinity (Kd ∼10 nM). The methodology developed in this work will be useful for studying the mechanism of DNA binding by other multi-step intercalating ligands and proteins.  相似文献   

20.
Physical origin of DNA condensation by multivalent cations remains unsettled. Here, we report quantitative studies of how one DNA-condensing ion (Cobalt3+ Hexammine, or Co3+Hex) and one nonDNA-condensing ion (Mg2+) compete within the interstitial space in spontaneously condensed DNA arrays. As the ion concentrations in the bath solution are systematically varied, the ion contents and DNA-DNA spacings of the DNA arrays are determined by atomic emission spectroscopy and x-ray diffraction, respectively. To gain quantitative insights, we first compare the experimentally determined ion contents with predictions from exact numerical calculations based on nonlinear Poisson-Boltzmann equations. Such calculations are shown to significantly underestimate the number of Co3+Hex ions, consistent with the deficiencies of nonlinear Poisson-Boltzmann approaches in describing multivalent cations. Upon increasing the concentration of Mg2+, the Co3+Hex-condensed DNA array expands and eventually redissolves as a result of ion competition weakening DNA-DNA attraction. Although the DNA-DNA spacing depends on both Mg2+ and Co3+Hex concentrations in the bath solution, it is observed that the spacing is largely determined by a single parameter of the DNA array, the fraction of DNA charges neutralized by Co3+Hex. It is also observed that only ∼20% DNA charge neutralization by Co3+Hex is necessary for spontaneous DNA condensation. We then show that the bath ion conditions can be reduced to one variable with a simplistic ion binding model, which is able to describe the variations of both ion contents and DNA-DNA spacings reasonably well. Finally, we discuss the implications on the nature of interstitial ions and cation-mediated DNA-DNA interactions.  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