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.
We have studied the NaCl-induced condensation of calf thymus chromatin by static light scattering of 90 degrees and shown that the increase in NaCl concentration up to 120-170 mM results in a large increase in scattering intensity of the total chromatin. Histones H1-depleted and trypsinized chromatin preparations do not reveal such a large increase in scattering intensity. The increase in the scattering intensity reflects the folding of the chromatin filaments, but not their aggregation. We have used this effect to monitor the kinetics of the chromatin condensation in response to a jump to higher NaCl concentrations by means of a stopped-flow technique. The results show that the condensation is a fast complex process consisting of at least two steps. The first step is only partially resolved by the stopped-flow apparatus. The second step has a time constant in the range of 20-50 ms, which does not depend on chromatin concentration.  相似文献   

2.
Abstract

We have studied the NaCl-induced condensation of calf thymus chromatin by static light scattering of 90° and shown that the increase in NaCl concentration up to 120–170 mM results in a large increase in scattering intensity of the total chromatin. Histones H1-depleted and trypsinized chromatin preparations do not reveal such a large increase in scattering intensity. The increase in the scattering intensity reflects the folding of the chromatin filaments, but not their aggregation. We have used this effect to monitor the kinetics of the chromatin condensation in response to a jump to higher NaCl concnetrations by means of a stopped-flow technique. The results show that the condensation is a fast complex process consisting of at least two steps. The first step is only partially resolved by the stopped-flow apparatus. The second step has a time constant in the range of 20–50 ms, which does not depend on chromatin concentration.  相似文献   

3.
Using isopycnic centrifugation in metrizamide gradient, effect of ions and chromatin concentration on the buoyant density of chromatin was quantitatively examined. An elevation followed by gradual decline and secondary increase of the density occurred in accordance with increase in MgCl2 or NaCl concentration. Maximum density was observed at a concentration of these salts known to result in the condensation of chromatin. Release of protein occurred during the phase of density decline. The second increase in density is mainly due to the density increment of DNA in the chromatin. The density was dependent upon the concentration of chromatin in a band formed in the metrizamide gradient, while the density of free DNA and protein was not so greatly affected by their concentration. The density of chromatin in the presence of 0.14 M NaCl was less affected by the chromatin concentration than that in the absence of salt. Calculation of results indicates that grade of hydration of chromatin at concentrations lower than 400 microgram/ml in 1 mM Tris-HCl (pH 8.0) is higher than that expected from its DNA and protein components.  相似文献   

4.
We present here a systematic study of ionic strength and divalent cation effects on Vinca alkaloid-induced tubulin spiral formation. We used sedimentation velocity experiments and quantitative fitting of weight-average sedimentation coefficients versus free drug concentrations to obtain thermodynamic parameters under various solution conditions. The addition of 50-150 mM NaCl to our standard buffer (10 mM piperazine-N,N'-bis(2-ethanesulfonic acid), 1 mM Mg, 50 microM GDP or GTP, pH 6.9) enhances overall vinblastine- or vincristine-induced tubulin self-association. As demonstrated in previous studies, GDP enhances overall self-association more than GTP, although in the presence of salt, GDP enhancement is reduced. For example, in 150 mM NaCl, GDP enhancement is 0.24 kcal/mol for vinblastine and 0.36 kcal/mol for vincristine versus an average enhancement of 0.87 (+/- 0.34) kcal/mol for the same drugs in the absence of salt. Wyman linkage analysis of experiments with vinblastine or vincristine over a range of NaCl concentrations showed a twofold increase in the change in NaCl bound to drug-induced spirals in the presence of GTP compared to GDP. These data indicate that GDP enhancement of Vinca alkaloid-induced tubulin self-association is due in part to electrostatic inhibition in the GTP state. In the absence of NaCl, we found that vinblastine and 1 mM Mn2+ or Ca2+ causes immediate condensation of tubulin. The predominant aggregates observed by electron microscopy are large sheets. This effect was not found with 1 mM Mg2+. At 100 microM cation concentrations (Mn2+, Mg2+, or Ca2+), GDP enhances vinblastine-induced spiral formation by 0.55 (+/- 0.26) kcal/mol. This effect is found only in K2, the association of liganded heterodimers at the ends of growing spirals. There is no GDP enhancement of K1, the binding of drug to heterodimer, although K1 is dependent upon the divalent cation concentration. NaCl diminishes tubulin condensation, probably by inhibiting lateral association, and allows an investigation of higher divalent cation concentrations. In the presence of 150 mM NaCl plus 1 mM divalent cations (Mn2+, Mg2+, or Ca2+) GDP enhances vinblastine-induced spiral formation by 0.35 (+/- 0.21) kcal/mol. Relaxation times determined by stopped-flow light scattering experiments in the presence of 150 mM NaCl and vincristine are severalfold longer than those in the presence of vinblastine, consistent with a mechanism involving the redistribution of longer polymers. Unlike previous results in the absence of NaCl, relaxation times in the presence of NaCl are only weekly protein concentration dependent, suggesting the absence of annealing or an additional rate-limiting step in the mechanism.  相似文献   

5.
In order to interpret the change in the X-ray scattering profiles from rat thymus chromatin, extensive model calculation was carried out. Chromatin is modelled as a string of subunits (nucleosomes) in which disorder is introduced into the positions of adjacent subunits. Disposition parameters characterizing the arrangement of subunits were estimated for various states of chromatin, so that the main feature of the scattering profiles is described. The result indicated that the structure of chromatin changes, as the NaCl concentration increases, from the extended "beads-on-a string" structure to the condensed helical structure. The latter has an outer diameter of about 26 nm with 3-4 nucleosomes per turn. In the intermediate state, it has a loose helical structure. The estimation of disorder suggested that the arrangement of subunits is appreciably disordered even in the condensed helical filament at 50 mM NaCl. Our model for chromatin condensation seems to support models of the "crossed linker" type.  相似文献   

6.
The optical anisotropy of chromatin with different length of the linker DNA isolated from a variety of sources (Frend erythroleukemia cells, calf thymus, hen erythrocytes and sea urchin sperm) has been studied in a large range of mono- and bivalent cations concentrations by the use of flow linear dichroism (LD) and electric dichroism. We have found that all chromatins studied displayed negative LD values in the range of 0.25 mM EDTA - 2 mM NaCl and close positive values in the range of 2-100 mM NaCl. Mg2+ cations, in contrast to Na+ cations, induce optically isotropic chromatin fibers. All chromatin samples exhibit positive form effect amounting to 5-10% of LD amplitude observed at 260 nm. This form effect is determined by the anisotropic scattering of polarized light by single chromatin fibers. The conformational transition at 2 mM NaCl leads to the distortion of chromatin filament structure. The reversibility of this distortion depends on the length of the linker DNA - for chromatins with the linker DNA of 10-30 b.p. it is parially reversible, while for preparations with longer linker DNA it is irreversible. Relatively low electric field does not affect chromatin structure, while higher electric field (more than 7 kV/cm) distorts the structure of chromatin. Presented results explain the contradictory data obtained by electrooptical and hydrooptical methods.  相似文献   

7.
Salt induces aggregation of large chromatin fragments maximally at 150-200 mM NaCl. The soluble fragments are depleted of H1 histones while the aggregated fragments are enriched. H1 histones did not equilibrate between the soluble and insoluble chromatin fractions when they were recycled through the process of salt-induced aggregation. The chromatin fragments that resisted aggregation retained more H1c subtype than they did H1 ab, correlating with previous results which showed complexes of H1c with DNA resisted salt-induced aggregation much more than complexes of DNA with other subtypes. The chromatin that was soluble at physiological concentrations of NaCl was DNase I sensitive and enriched in acetylated core histones. We conclude that H1 histone is nonuniformly distributed in chromatin in a stable pattern that probably correlates with the different degrees of condensation known to exist in vivo.  相似文献   

8.
The optical anisotropy of chromatin with different length of the linker DNA isolated from a variety of sources (Friend erythroleukemia cells, calf thymus, hen erythrocytes and sea urchin sperm) has been studied in a large range of mono- and bivalent cations by the use of flow linear dichroism and electric dichroism. We have found that all chromatins studied displayed negative LD values in the range of 0.25 mM EDTA--2 mM NaCl and close positive values in the range of 2-100 mM NaCl. Mg2+ cations, in contrast to Na+ cations, induce optically isotropic chromatin fibers. All chromatin samples exhibit positive form effect amounting to 5-10% of LD amplitude observed at 260 nm. This form effect is determined by the anisotropic scattering of polarized light by single chromatin fibers. The conformational transition at 2 mM NaCl leads to the distortion of chromatin filament structure. The reversibility of this distortion depends on the length of the linker DNA--for chromatins with linker DNA 10-30 b.p. it is partially reversible, while for preparations with longer linker DNA it is irreversible. Relatively low electric fields do not have an effect on chromatin structure, while higher electric fields (more than 7 kV/cm) distort the structure of chromatin.  相似文献   

9.
Comparison between the internucleosomal distance found by X-ray solution scattering for chicken erythrocyte (23 nm) and sea urchin (30 nm) chromatin indicates that this distance is proportional to the linker length. The diameter of the condensed sea urchin chromatin fibers is about 45 nm which is significantly larger than in chicken erythrocyte chromatin (35 nm). Trivalent cations (Gd, Tb, Cr) and the polyamines spermine and spermidine were found to induce compaction at much lower concentrations than the divalent cations but Gd, Tb and Cr induce aggregation before full compaction of the fibers. The influence of hydrogen bonding is illustrated by comparison of the effects of NaCl, ammonium chloride and alkylammonium chlorides on condensation. Solubility experiments indicate that there is a nearly linear dependence of the Mg-- concentration at which precipitation occures on chromatin concentration and confirm the differences between cations observed by X-ray scattering.The chicken erythrocyte chromatin samples were further characterized by their reduced electric dichroism. The values found are consistent with the model derived from X-ray scattering and are compared with those reported in the literature.  相似文献   

10.
The condensation of the 10 nm chromatin filament in the 30 nm fiber by monovalent cations, polyamines and bivalent cations was studied with light scattering at 90 degrees and flow linear dichroism methods. It was found that monovalent cation- and polyamine-induced folding was a two-step process: a precondensation, when a rotation of nucleosomes takes place only, and a condensation step without changes in nucleosome orientation. Divalent cations affected the structure of chromatin in one step only -- condensation of the chromatin filament being accompanied by nucleosome reorientation.  相似文献   

11.
Using small-angle neutron scattering (SANS), we have measured the salt-dependent static structure factor of di- and trinucleosomes from chicken erythrocytes and from COS-7 cells. We also determined the sedimentation coefficients of these dinucleosomes and dinucleosomes reconstituted on a 416-bp DNA containing two nucleosome positioning sequences of the 5S rDNA of Lytechinus variegatus at low and high salt concentrations. The internucleosomal distance d was calculated by simulation as well as Fourier back-transformation of the SANS curves and by hydrodynamic simulation of sedimentation coefficients. Nucleosome dimers from chicken erythrocyte chromatin show a decrease in d from approximately 220 A at 5 mM NaCl to 150 A at 100 mM NaCl. For dinucleosomes from COS-7 chromatin, d decreases from 180 A at 5 mM to 140 A at 100 mM NaCl concentration. Our measurements on trinucleosomes are compatible with a compaction through two different mechanisms, depending on the salt concentration. Between 0 and 20 mM NaCl, the internucleosomal distance between adjacent nucleosomes remains constant, whereas the angle of the DNA strands entering and leaving the central nucleosome decreases. Above 20 mM NaCl, the adjacent nucleosomes approach each other, similar to the compaction of dinucleosomes. The internucleosomal distance of 140-150 A at 100 mM NaCl is in agreement with distances measured by scanning force microscopy and electron microscopy on long chromatin filaments.  相似文献   

12.
The chromatin structure in solution has been studied by the flow linear dichroism method (LD) in a wide range of ionic strengths. It is found that increasing the ionic strength from 0.25 mM Na2EDTA, pH 7.0 to 100 mM NaCl leads to a strong reduction of the LD amplitude of chromatin and inversion of the LD sign from negative to positive at 2 mM NaCl. Chromatin exhibits a positive LD maximum value at 10-20 mM NaCl. These data enable us to conclude that in very low ionic strength (0.25 mM Na2EDTA) the nucleosome discs are oriented with their flat faces more or less parallel to the chromatin filament axis. Increasing ionic strength up to 20 mM NaCl leads to reorientation of the nucleosome discs and to formation of chromatin structures with nucleosome flat faces inclined to the fibril axis. A conformational transition of that kind is not revealed in H1-depleted chromatin. The condensation of the chromatin filaments with increasing concentration of NaCl from 20 mM to 100 mM slightly influences the orientation of the nucleosomes.  相似文献   

13.
Linker histones are involved in chromatin higher-order structure and gene regulation. We have successfully achieved partial phosphorylation of linker histones in chicken erythrocyte soluble chromatin with CDK2, as indicated by HPCE, MALDI-TOF and Tandem MS. We have studied the effects of linker histone partial phosphorylation on secondary structure and chromatin condensation. Infrared spectroscopy analysis showed a gradual increase of β-structure in the phosphorylated samples, concomitant to a decrease in α-helix/turns, with increasing linker histone phosphorylation. This conformational change could act as the first step in the phosphorylation-induced effects on chromatin condensation. A decrease of the sedimentation rate through sucrose gradients of the phosphorylated samples was observed, indicating a global relaxation of the 30-nm fiber following linker histone phosphorylation. Analysis of specific genes, combining nuclease digestion and qPCR, showed that phosphorylated samples were more accessible than unphosphorylated samples, suggesting local chromatin relaxation. Chromatin aggregation was induced by MgCl2 and analyzed by dynamic light scattering (DLS). Phosphorylated chromatin had lower percentages in volume of aggregated molecules and the aggregates had smaller hydrodynamic diameter than unphosphorylated chromatin, indicating that linker histone phosphorylation impaired chromatin aggregation. These findings provide new insights into the effects of linker histone phosphorylation in chromatin condensation.  相似文献   

14.
The superstructure of chromatin and its condensation mechanism   总被引:8,自引:0,他引:8  
Synchroton radiation X-ray scattering experiments have been performed on chicken erythrocyte chromatin fibres over a wide range of ionic conditions and on various states of the fibres (i.e. "native" in solution, in gels and in whole nuclei; chromatin depleted of the H1 (H5) histones and chromatin with bound ethidium bromide). A correlation between the results obtained with the various chromatin preparations provides evidence for a model according to which at low ionic strength the chromatin fibre already possesses a helical superstructure, with a diameter comparable to that of condensed chromatin, held together by the H1(H5) histone. The most significant structural modification undergone upon an increase of the ionic strength is a reduction of the helix pitch, this leads to condensation in a manner similar to the folding of an accordion. The details of this process depend on whether monovalent or divalent cations are used to raise the ionic strength, the latter producing a much higher degree of condensation. Measurements of the relative increase of the mass per unit length indicate that the most condensed state is a helical structure with a pitch around 3.0-4.0 nm. In this paper we give a detailed presentation of the experimental evidence obtained from static and time-resolved scattering experiments, which led to this model.  相似文献   

15.
Calf thymus chromatin gel, containing strongly bound nonhistone proteins, was used to study the effect of easily removable and tightly bound cations on the condensation of chromatin. The chromatin volume was found to be linearly dependent on the reciprocal square root of the concentration of easily removable cations (Tris X H+ + Na+ and Mg2+) except for the initial stages of condensation (up to 7-10 mM monovalent and 0.15-0.2 mM divalent cations). The effect of Mg2+ at the initial stage of condensation was not reproduced by Na+ and vice versa. At higher concentrations the effects of Na+ and Mg2+ were additive. The removal of tightly bound divalent cations by a treatment of the chromatin gel with 1,10-phenanthroline led to an approx. 50% increase in the volume of the chromatin gel, which was maintained at each concentration of easily removable cations. The 1,10-phenanthroline-caused decondensation of the chromatin gel was reversed by Ca2+ but not by Mg2+, Zn2+ and Cu2+. The chromatin gel pretreated with Ca2+ was not further decondensed by 1,10-phenanthroline.  相似文献   

16.
Summary. Leaf senescence is a highly coordinated process which involves programmed cell death (PCD). Early stages of leaf senescence occurring during normal leaf ontogenesis, but not triggered by stress factors, are less well known. In this study, we correlated condensation of chromatin and nuclear DNA (nDNA) fragmentation, two main features of PCD during early senescence in barley leaves, with the appearance of nitric oxide (NO) within leaf tissue. With the help of the alkaline version of the comet assay, together with measurements of nDNA fluorescence intensity, we performed a detailed analysis of the degree of nDNA fragmentation. We localised NO in vivo and in situ within the leaf and photometrically measured its concentration with the NO-specific fluorochrome 4-amino-5-methylamino-2′,7′-difluorofluorescein. We found that both nDNA fragmentation and chromatin condensation occurred quite early during barley leaf senescence and always in the same order: first nDNA fragmentation, in leaves of 6-day-old seedlings, and later chromatin condensation, in the apical part of leaves from 10-day-old seedlings. PCD did not start simultaneously even in neighbouring cells and probably did not proceed at the same rate. NO was localised in vivo and in situ within the cytoplasm, mainly in mitochondria, in leaves at the same stage as those in which chromatin condensation was observed. Localisation of NO in vascular tissue and in a large number of mesophyll cells during the senescence process might imply its transport to other parts of the leaf and its involvement in signalling between cells. The fact that the highest concentration of NO was found in the cytoplasm of mesophyll cells in the earliest stage of senescence and lower concentrations were found during later stages might suggest that NO plays an inductive role in PCD. Correspondence: A. Mostowska, Department of Plant Anatomy and Cytology, Institute of Experimental Biology of Plants, University of Warsaw, Miecznikowa 1, 02-096 Warsaw, Poland.  相似文献   

17.
Trypanosoma cruzi is the etiological agent of Chagas. Although the nuclear chromatin of this parasite is organized in the form of nucleosome filaments, its chromatin is physically and enzymatically fragile, and no condensation into chromosomes occurs during mitosis. All previous investigations have been carried out with epimastigote form in its proliferate stage. It is not known whether these differences in chromatin structure are also found in the non-proliferate stationary epimastigote forms and in tissue derived trypomastigotes. Our results confirm that chromatin of logarithmic epimastigotes presents limited compaction when increasing salt concentrations from 1 to 100 mM NaCl, and no 30-nm fibers were formed. Contrary to these results, non-proliferative forms of the parasites showed a pattern of compactation similar to that observed in rat liver chromatin, where solenoids of 30-nm fibers are formed at 100-mM NaCl. In accordance with these results, digestion of the nuclear chromatin with DNase I revealed that the chromatin of logarithmic phase epimastigotes was more accessible to the enzyme. We conclude from these results that structural differences in the chromatin exist not only between T. cruzi and higher eukaryotes but also among various forms of the parasite. The functional significance of these differences are currently under investigation.  相似文献   

18.
Kinetics of cooperative binding of rat polymerase beta to a double-stranded DNA has been studied using the fluorescence stopped-flow techniques. The data have been analyzed by an approach developed to examine complete kinetics of cooperative large ligand binding to a one-dimensional lattice. The method is based on using the smallest possible system that preserves key ingredients of cooperative binding; i.e., at saturation, the lattice can accept only two ligand molecules. It allows the identification of collective amplitudes as well as amplitudes describing particular normal modes of the reaction. The mechanism of the intrinsic binding of pol beta to the dsDNA is different from the analogous mechanism for the ssDNA. The difference originates from different enzyme orientations in the corresponding complexes. Intrinsic binding to the dsDNA includes only two sequential steps: a very fast bimolecular association followed by an energetically favorable conformational transition of the complex. The transition following the bimolecular step does not facilitate the engagement of the enzyme in cooperative interactions. Its role seems to be reinforcing the affinity of the bimolecular step. Salt and magnesium cations affect both the bimolecular step and the conformational transition. As a result, the bimolecular step is less sensitive to the increased salt concentration, allowing the enzyme to preserve its initial dsDNA affinity. The changing character of cooperative interactions between bound enzyme molecules as a function of NaCl concentration and MgCl(2) concentration does not affect the binding mechanism. The engagement in cooperative interactions is approximately 3-4 orders of magnitude slower than the conformational transition of the DNA-bound polymerase. The importance of the obtained results for the pol beta activities is discussed.  相似文献   

19.
M Kubista  T H?rd  P E Nielsen  B Nordén 《Biochemistry》1985,24(23):6336-6342
We have studied the structure of nuclease-solubilized chromatin from Ehrlich ascites cells by flow linear dichroism (LD) using the anisotropic absorption of the DNA bases and of two intercalated dyes, ethidium bromide and methylene blue. It is confirmed that intercalation occurs preferentially in the linker part of the chromatin fiber, at binding ratios (dye/base) below 0.020. Using this information, we determined the orientation of the linker in relation to the average DNA organization in chromatin. The LD measurements indicate that the conformation of chromatin is considerably changed in the ionic strength interval 0.1-10 mM NaCl: with increasing salt concentration, the LD of the intrinsic DNA base absorption changes signs, from negative to positive, at approximately 2.5 mM NaCl. The LD of the intercalated dyes also changes signs, however, at a somewhat higher salt concentration. The results are analyzed in terms of possible allowed combinations of tilt angles of nucleosomes and pitch or tilt angles of linker DNA sections relative to the fiber axis, at different salt concentrations in the interval 0.1-10 mM NaCl. Two models for the salt-induced structural change of chromatin are discussed.  相似文献   

20.
Polyamine-DNA interactions. Condensation of chromatin and naked DNA   总被引:2,自引:0,他引:2  
We have used flow linear dichroism (LD) and light scattering at 90 degrees to study the condensation of both DNA and calf thymus chromatin by polyamines, such as spermine, spermidine and its analogs designated by formula NH3+(CH2)iNH2+(CH2)jNH3+, where i = 2,3 and j = 2,3, putrescine, cadaverine and MgCl2. It has been found that the different polyamines affect DNA and chromatin in a similar way. The level of compaction of the chromatin fibers induced by spermine, spermidine and the triamines NH3+(CH2)3NH2+(CH2)3NH3+ and NH3+(CH2)3NH2+(CH2)2NH3+ and MgCl2 is found to be identical. The triamine NH3+(CH2)3NH2+(CH2)2NH3+ and the diamines studied condense neither chromatin nor DNA. This drastic difference in the action of the triamines indicates that not only the charge, but also the structure of the polycations might play essential roles in their interactions with DNA and chromatin. It is shown that a mixture of mono- and multivalent cations affect DNA and chromatin condensation competitively, but not synergistically, as claimed in a recent report by Sen and Crothers (Biochemistry 25, 1495-1503, 1986). We have also estimated the extent of negative charge neutralization produced by some of the polyamines on their binding to chromatin fibers. The stoichiometry of polyamine binding at which condensation of chromatin is completed is found to be two polyamine molecules per DNA turn. The extent of neutralization of the DNA phosphates by the histones in these compact fibers is estimated to be about 55%. The model of polyamine interaction with chromatin is discussed.  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