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.
Teif  V. B.  Lando  D. Yu. 《Molecular Biology》2001,35(1):106-107
A method for calculating the curves of DNA transition from linear to condensed state upon binding of condensing ligands has been developed. The character of the transition and ligand concentration necessary for condensation have been shown to be governed by the length of DNA molecule, energy and stoichiometry parameters of the DNA–ligand complex (equilibrium constant between linear and condensed form in the absence of ligands, constants for ligand binding to linear and condensed forms, the number of base pairs covered by one ligand, etc.). The results of the calculations indicate that a slight difference in the free energies of these DNA states (less than 6 cal/mol(bp) for a DNA of 500 bp) is sufficient for the existence of a stable linear state in the absence of ligands (in free DNA) and the formation of stable condensed state upon complexation.  相似文献   

2.
A method for calculating the curves of DNA transition from linear to condensed state upon binding of condensing ligands has been developed. The character of the transition and ligand concentration necessary for condensation have been shown to be governed by the length of DNA molecule, energy and stoichiometry parameters of DNA-ligand complex (equilibrium constant between linear and condensed form in the absence of ligands, constants for ligand binding to linear and condensed forms, the number of base pairs covered by one ligand, etc.). The results of the calculations indicate that only slight difference in the free energies of these states in free DNA (less than 6 cal/mole(bp) for DNA of 500 bp long) is sufficient for the existence of stable linear state in the absence of ligands (in free DNA) and the formation of stable condensed state upon complexation.  相似文献   

3.
Teif VB 《Biophysical journal》2005,89(4):2574-2587
We test and compare different models for ligand-induced DNA condensation. Using 14C-labeled spermidine3+, we measure the binding to condensed DNA at micromolar to molar polyamine concentrations. DNA aggregates at a critical polyamine concentration. Spermidine3+ binding becomes highly cooperative at the onset of aggregation. At higher concentrations, spermidine3+ binding to condensed DNA reaches a plateau with the degree of binding equal to 0.7 (NH(4+)/PO3-). Condensed DNA exists in a wide range of spermidine concentrations with the roughly constant degree of ligand binding. At greater concentrations, the degree of binding increases again. Further spermidine penetration between the double helices causes DNA resolubilization. We show that a simple two-state model without ligand-ligand interactions qualitatively predicts the reentrant aggregation-resolubilization behavior and the dependence on the ligand, Na+, and DNA concentrations. However, such models are inconsistent with the cooperative ligand binding to condensed DNA. Including the contact or long-range ligand-ligand interactions improves the coincidence with the experiments, if binding to condensed DNA is slightly more cooperative than to the starting DNA. For example, in the contact interaction model it is equivalent to an additional McGhee-von Hippel cooperativity parameter of approximately 2. Possible physical mechanisms for the observed cooperativity of ligand binding are discussed.  相似文献   

4.
Long-range interaction between all the ligands bound to DNA molecule may give rise to adsorption with the character of phase transition of the first kind (D. Y. Lando, V. B. Teif, J. Biomol. Struct & Dynam. 18, 903-911 (2000)). In this case, the binding curve, c(c(o)), is characterized by a sudden change of the relative concentration of bound ligands ((c)) at a critical concentration of free (unbound) ligands, c(o)=c(ocr), from a low c value to a high one where c(o) is molar concentration of free ligands. Such a transition might be caused by some types of DNA condensation or changes in DNA topology. For the study of the conditions necessary for adsorption with the character of phase transition, a calculation procedure based on the method of the free energy minimum is developed. The ligand size and two types of interactions between ligands adsorbed on DNA molecule are taken into consideration: long-range interaction between all the ligands bound to DNA and contact interactions between neighboring ligands. It was found that a) Stronger long-range interaction is required for longer ligands to induce phase transition that is occurred at greater c(ocr) values; b) Pure contact interaction between neighboring ligands can not itself initiate phase transition. However contact cooperativity strongly decreases the threshold value of energy of long-range interaction necessary to give rise to the transition.  相似文献   

5.
The importance of understanding the dynamics of DNA condensation is inherent in the biological significance of DNA packaging in cell nuclei, as well as for gene therapy applications. Specifically, the role of ligand hydrophobicity in DNA condensation has received little attention. Considering that only multivalent cations can induce true DNA condensation, previous studies exploring monovalent lipids have been unable to address this question. In this study we have elucidated the contribution of the hydrophobic effect to multivalent cation- and cationic lipid-DNA binding and DNA collapse by studying the thermodynamics of cobalt hexammine-, spermine-, and lipospermine-plasmid DNA binding at different temperatures. Comparable molar heat capacity changes (DeltaC(p)) associated with cobalt hexammine- and spermine-DNA binding (-23.39 cal/mol K and -17.98 cal/mol K, respectively) suggest that upon binding to DNA, there are insignificant changes in the hydration state of the methylene groups in spermine. In contrast, the acyl chain contribution to the DeltaC(p) of lipospermine-DNA binding (DeltaC(p ) = DeltaC(p lipospermine) - DeltaC(p spermine)) is significant (-220.94 cal/mol K). Although lipopermine induces DNA ordering into "tubular" suprastructures, such structures do not assume toroidal dimensions as observed for spermine-DNA complexes. We postulate that a steric barrier posed by the acyl chains in lipospermine precludes packaging of DNA into dimensions comparable to those found in nature.  相似文献   

6.
Sarkar R  Pal SK 《Biopolymers》2006,83(6):675-686
We have studied intercalation of ethidium bromide (EB) to genomic DNA encapsulated in a nanospace of an anionic AOT reverse micelle (RM). Circular dichroism (CD) study on the DNA in the RM reveals its condensed form. Here, we have used temporal decay-associated spectra (DAS) and time-resolved area normalized emission spectral (TRANES) techniques to investigate EB-binding to condensed DNA because the interference of emission from unbound EB in the RM makes conventional steady state and picosecond resolved fluorescence spectroscopic techniques challenging. The binding affinity of the ligand EB with the DNA in the RM is found to increase with the size of the RM, reflecting the effect of lessening of DNA condensation on the binding affinity. CD spectra of the DNA in the RM with various sizes indicate the structural change of the condensed DNA with reverse micellar size. DAS and TRANES techniques along with dynamic light scattering studies of the EB-DNA complex in the RM further reveal two kinds of binding modes of the ligand with the condensed DNA even in essentially monodispersed RMs. To investigate the role of RM on the ligand binding and secondary structure of the DNA, we have also studied complexation of EB with two synthetic self-complimentary oligonucleotides of sequences (CGCAAATTTGCG)2 and (CGCGCGCGCGCG)2 in the RM.  相似文献   

7.
Influence of long-range interactions between ligands bound to DNA molecule on the character of their adsorption is studied using computer modeling. For this investigation, two calculation procedures are developed. They are based upon the method of the free energy minimum and on the partition function method. The both procedures demonstrate that in the case of a strong enough attraction between all the bound ligands their binding to DNA has the character of phase transition of the first kind. There is a break in the binding curve c(c0) where c - relative concentration of bound ligands, c0 - molar concentration of free ligands. The break occurs because there is an interval of central degrees of binding (approximately 50% of the maximum c value) that are prohibited for individual DNA molecules. Such a transition might be caused by some types of DNA condensation. Attraction between the neighboring ligands only, adjacent or/and separated by double helix regions, does not cause this effect.  相似文献   

8.
We have evaluated the effect of NaCl concentration on the mode of binding of poly-L-lysine to DNA and the resulting structural and functional features of the condensed DNA particles using DNA precipitation, DNase I resistance, electron microscopy, and receptor-mediated gene transfer assays. At a high concentration of NaCl and in the presence of excess DNA, poly-L-lysine interacted with DNA cooperatively, fully condensing some of the DNA and leaving the rest of the DNA unbound. At low NaCl concentrations, poly-L-lysine molecules interacted with DNA in a noncooperative fashion, i.e. they bind randomly to the whole population of DNA molecules. Cooperative binding of poly-L-lysine to DNA occurred over a narrow range of NaCl concentrations, and the specific salt concentration depended on the length of the poly-L-lysine. The ability of condensed DNA to withstand digestion by DNase I was correlated with the structural features of the condensed DNA as determined by electron microscopy. Using our condensation procedure, cooperative binding of poly-L-lysine to DNA is a necessary prerequisite for the preparation of condensed DNA having a spherical shape and a diameter of 15-30 nm. Condensed DNA, containing galactosylated poly-L-lysine, was evaluated further for the extent and specificity of receptor-mediated gene transfer into HuH-7 human hepatoma cells via the asialoglycoprotein receptor. Efficient receptor-mediated transfection occurred only when condensed DNA complexes had a spherical shape with a diameter of 15-30 nm; asialofetuin, a natural ligand for the asialoglycoprotein receptor, inhibited this process by up to 90%. Our results support the importance of appropriate DNA condensation for the uptake and ultimate expression of DNA in hepatic cells.  相似文献   

9.
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   

10.
St Maurice M  Bearne SL 《Biochemistry》2004,43(9):2524-2532
Mandelate racemase (EC 5.1.2.2) from Pseudomonas putida catalyzes the interconversion of the two enantiomers of mandelic acid with remarkable proficiency, stabilizing the altered substrate in the transition state by approximately 26 kcal/mol. We have used a series of substrate analogues (glycolates) and intermediate analogues (hydroxamates) to evaluate the contribution of the hydrophobic cavity within the enzyme's active site to ligand binding. Free energy changes accompanying binding of glycolate derivatives correlated well with the hydrophobic substituent constant pi and the van der Waals surface areas of the ligands. The observed dependence of the apparent binding free energy on surface area of the ligand was -30 +/- 5 cal mol(-1) A(-2) at 25 degrees C. Free energy changes accompanying binding of hydroxamate derivatives also correlated well with pi values and the van der Waals surface areas of the ligands, giving a slightly greater free energy dependence equal to -41 +/- 3 cal mol(-1) A(-2) at 25 degrees C. Surprisingly, mandelate racemase exhibited a binding affinity for the intermediate analogue benzohydroxamate that was 2 orders of magnitude greater than that predicted solely on the basis of hydrophobic interactions. This suggests that there are additional specific interactions that stabilize the altered substrate in the transition state. Mandelate racemase was competitively inhibited by (R,S)-1-naphthylglycolate (apparent K(i) = 1.9 +/- 0.1 mM) and (R,S)-2-naphthylglycolate (apparent K(i) = 0.52 +/- 0.03 mM), demonstrating the plasticity of the hydrophobic pocket. Both (R)- (K(m) = 0.46 +/- 0.06 mM, k(cat) = 33 +/- 1 s(-1)) and (S)-2-naphthylglycolate (K(m) = 0.41 +/- 0.03 mM, k(cat) = 25 +/- 1 s(-1)) were shown to be alternative substrates for mandelate racemase. These kinetic results demonstrate that no major steric restrictions are imposed on the binding of this bulkier substrate in the ground state but that steric factors appear to impair transition state/intermediate stabilization. 2-Naphthohydroxamate was identified as a competitive inhibitor of mandelate racemase, binding with an affinity (K(i) = 57 +/- 18 microM) that was reduced relative to that observed for benzohydroxamate and that was in accord with the approximately 10-fold reduction in the value of k(cat)/K(m) for the racemization of 2-naphthylglycolate. These findings indicate that, for mandelate racemase, steric constraints within the hydrophobic cavity of the enzyme-intermediate complex are more stringent than those in the enzyme-substrate complex.  相似文献   

11.
In studies aimed at developing a high affinity IL-8 antagonist, our first objective was to generate a high-affinity IL-8 analogue. We targeted amino acids within the receptor-binding domain and found that IL-8((3-73))K11R induced significantly more neutrophil beta-glucuronidase release than either IL-8 or the alternate analogues and, in chemotaxis assays, induced 2-3-fold greater neutrophil responses than IL-8. Furthermore, in competitive radio- or biotinylated-ligand binding assays, IL-8((3-73))K11R was more effective than IL-8, IL-8((3-73)), or its T12S, H13F, and K11R/T12S/H13F analogues in blocking IL-8 binding to neutrophils; 1.8 pmol IL-8((3-73))K11R inhibited by 50% the binding of approximately 20 pmol (125)I-IL-8 to neutrophils. Both IL-8 (a CXCR1/CXCR2 ligand) and the CXCR2-specific ligand GROalpha differentially inhibited binding of (125)I-IL-8((3-73))K11R to neutrophils, albeit weakly, suggesting that IL-8((3-73))K11R is a high affinity ligand for both the CXCR1 and CXCR2. Thus IL-8((3-73))K11R is an excellent candidate for further studies aimed at generating a high affinity IL-8 antagonist.  相似文献   

12.
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   

13.
Dynamics of DNA condensation   总被引:6,自引:0,他引:6  
D Porschke 《Biochemistry》1984,23(21):4821-4828
The condensation of DNA induced by spermine and spermidine is investigated by equilibrium titrations and stopped-flow and field-jump experiments using scattered light detection. The spermine concentration required for the cooperative condensation process is measured at different DNA concentrations; these data are used to evaluate both the condensation threshold degree of spermine binding and the binding constant of spermine according to an excluded-site model. Stopped-flow measurements of the spermine-induced condensation demonstrate the existence of two processes: (1) A "fast" reaction is observed in the millisecond time range, when the reactant concentrations are around 1 microM; it is associated with a characteristic induction period and is assigned to the intramolecular condensation reaction. (2) A slow reaction with time constants of, e.g., 100 s strongly dependent upon both spermine and DNA concentrations is assigned to an intermolecular DNA association. The unusual time course of the intramolecular condensation reaction with the induction period provides evidence for a "threshold kinetics". During the induction period, spermine molecules are bound to DNA, but the degree of binding remains below the threshold value. As soon as the degree of ligand binding arrives at the threshold, the DNA is condensed in a relatively fast reaction. Model calculations of the spermine binding kinetics according to an excluded-site model demonstrate that the spermine molecules bound to DNA are mobile along the double helix. A comparison of the experimental data with the results of Monte Carlo simulations suggests a rate constant of approximately 200 s-1 for spermine movement by one nucleotide residue.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
Despite insights on the cellular level, the molecular details of chromatin reorganization in sperm development, which involves replacement of histone proteins by specialized factors to allow ultra most condensation of the genome, are not well understood. Protamines are dispensable for DNA condensation during Drosophila post-meiotic spermatogenesis. Therefore, we analyzed the interaction of Mst77F, another very basic testis-specific protein with chromatin and DNA as well as studied the molecular consequences of such binding. We show that Mst77F on its own causes severe chromatin and DNA aggregation. An intrinsically unstructured domain in the C-terminus of Mst77F binds DNA via electrostatic interaction. This binding results in structural reorganization of the domain, which induces interaction with an N-terminal region of the protein. Via putative cooperative effects Mst77F is induced to multimerize in this state causing DNA aggregation. In agreement, overexpression of Mst77F results in chromatin aggregation in fly sperm. Based on these findings we postulate that Mst77F is crucial for sperm development by giving rise to a unique condensed chromatin structure.  相似文献   

15.
16.
Summary The species-specific proportion of chromatin in the condensed state was estimated by quantitative electron microscopic morphometry of nuclear sections in 9 diploid and 5 allopolyploid species of Microseris (Asteraceae). A positive correlation between the genome size (haploid DNA content, or C value) and the percentage of chromatin in the condensed state (as visible in ultrathin sections) was found in diploids (r=0.89). Nuclei of allopolyploid (tetraploid) species exhibit condensed chromatin in a percentage which corresponds to the average of the values found in the parents. This suggests that each parental genome controls chromatin condensation at interphase independently within the nucleus, and that the degree of condensation is not directly determined by the nuclear DNA content per se. Genome size differences among Microseris species may depend preferentially, but not entirely, on DNA fractions located in, and perhaps being the cause of, condensed chromatin.Dedicated to Professor F. Mechelke in honour of his 60th birthday.  相似文献   

17.
To determine the effect of the redox state of the Rieske protein on ligand binding to the quinol oxidation site of the bc(1) complex, we measured the binding rate constants (k(1)) for stigmatellin and myxothiazol, at different concentrations of decylbenzoquinone or decylbenzoquinol, in the bovine bc(1) complex with the Rieske protein in the oxidized or reduced state. Stigmatellin and myxothiazol bound tightly and competitively with respect to quinone or quinol, independently of the redox state of the Rieske protein. In the oxidized bc(1) complex, the k(1) values for stigmatellin ( approximately 2.6 x 10(6) m(-1)s(-1)) and myxothiazol ( approximately 8 x 10(5) m(-1)s(-1)), and the dissociation constant (K(d)) for quinone, were similar between pH 6.5 and 9, indicating that ligand binding is independent of the protonation state of histidine 161 of the Rieske protein (pK(a) approximately 7.6). Reduction of the Rieske protein increased the k(1) value for stigmatellin and decreased the K(d) value for quinone by 50%, without modifying the k(1) for myxothiazol. These results indicate that reduction of the Rieske protein and protonation of histidine 161 do not induce a strong stabilization of ligand binding to the quinol oxidation site, as assumed in models that propose the existence of a highly stabilized semiquinone as a reaction intermediate during quinol oxidation.  相似文献   

18.
In the presence of cationic ligands, DNA molecules can become aggregated into larger particles in a process known as condensation. DNA condensates are of interest as models for the dense packing found in naturally occurring structures such as phage heads and chromatin. They have found extensive application in DNA transfection and also provide convenient models with which to study DNA damage by the direct effect of ionizing radiation. Further, conjugates of cationic peptides with fatty acids may represent a class of attractive ligands for these areas because of their simple synthesis. When plasmid pUC18 is used as the DNA target and N-caproyl-penta-arginine amide (Cap-R(5)-NH(2)) is used as the ligand, the physical properties of the resulting mixtures were characterized using static and dynamic light scattering, sedimentation, dye exclusion, circular dichroism, nanoparticle tracking, and atomic force microscopy. Their chemical properties were assayed using solvent extraction and protection against hydroxyl radical attack and nuclease digestion. Titration of the plasmid with the Cap-R(5)-NH(2) ligand produced sharply defined changes in both chemical and physical properties, which was associated with the formation of condensed DNA particles in the 100-2000 nm size range. The caproyl group at the ligand's N-terminus produced a large increase in the partitioning of the resulting condensate from water into chloroform and in its binding to the neutral detergent Pluronic F-127. Both the physical and chemical data were all consistent with condensation of the plasmid by the ligand where the presence in the ligand of the caproyl group conferred an extensive lipophilic character upon the condensate.  相似文献   

19.
Although structural differences for the Mg-DNA and Ca-DNA complexes are provided in the solid state, such comparative study in aqueous solution has been less investigated. The aim of this study was to examine the bindings of Mg and Ca cations with calf thymus DNA in aqueous solution at physiological pH, using constant concentration of DNA (1.25 or 12.5 mM) and various concentrations of metal ions (2 microM-650 microM). Capillary electrophoresis, UV-visible, and Fourier transform infrared spectroscopic methods were used to determine the cation-binding modes, the binding constants, and DNA structural variations in aqueous solution. Direct Ca-PO(2) binding was evident by major spectral changes (shifting and splitting) of the backbone PO(2) asymmetric stretching at 1222 cm(-1) with K = 4.80 x 10(5) M(-1), whereas an indirect Mg-phosphate interaction occurred (due to the lack of shifting and splitting of the phosphate band at 1222 cm(-1)) with K = 5.6 x 10(4) M(-1). The metal-base bindings were directly for the Mg with K = 3.20 x 10(5) M(-1) and indirectly for the Ca cation with K = 3.0 x 10(4) M(-1). Both major and minor groove bindings were observed with no alteration of the B-DNA conformation.  相似文献   

20.
Atomic force microscopy (AFM) has been used to visualize the process of condensation of plasmid DNA by poly-L-ornithine on mica surface. AFM images reveal that the transition of negatively charged DNA to condensed nanoparticles on addition of increasing amounts of positively charged poly-L-ornithine (charge ratio (Z+/Z-) varied between 0.1 and 1) at a wide range of DNA concentrations (3-20 ng/microl) occurs through formation of several distinct morphologies. The nature of the complexes is strongly dependent on both the charge ratio and the DNA concentration. Initiation of condensation when the concentration of DNA is low (approximately 3-7 ng/microl) occurs possibly through formation of monomolecular complexes which are thick rod-like in shape. On the contrary, when condensation is carried out at DNA concentrations of 13-20 ng/microl, multimolecular structures are also formed even at low charge ratios. This difference in pathway seems to result in differences in the extent of condensation as well as size and aggregation of the nanoparticles formed at the high charge ratios. To the best of our knowledge, this is the first direct single molecule elucidation of the mechanism of DNA condensation by poly-L-ornithine. Cationic poly-aminoacids like poly-L-ornithine are known to be efficient in delivery of plasmid DNA containing therapeutic genes in a variety of mammalian cell lines by forming condensed "nanocarriers" with DNA. Single molecule insight into the mechanism by which such nanocarriers are packaged during the condensation process could be helpful in predicting efficacy of intracellular delivery and release of DNA from them and also provide important inputs for design of new gene delivery vectors.  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