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 topology of the cell-cell contact seam formed when normal or pronase pre-treated (PPT) erythrocytes are exposed to wheat germ agglutinin (WGA) in isotonic media of different ionic strengths was examined here. Lectin uptake and cell agglutination were also quantified. Agglutination of normal cells was gradually and significantly inhibited as ionic strength (IS) was reduced from 0.15 (buffered 145 mm NaCl) to 0.105. Agglutination was less inhibited in PPT cells, even when IS was reduced to 0.09. Cell contact seams formed during agglutination showed patterns of localized contacts. The scale of the patterns, i.e. the average lateral separation distance of contact regions, was 0.62 microm for normal cells and was significantly shorter, at 0.44 microm, for PPT cells at an IS of 0.15. The scale increased significantly for both cell types when the IS was reduced to 0.09. Flow cytometry measurements showed that WGA uptake by normal cells increased slightly, whilst that for PPT cells was unchanged, as IS was decreased from 0.15 to 0.09. The results imply that, whilst ionic strength change does not exert a strong influence on intermolecular WGA-ligand binding, physico-chemical modification of the interaction between cells modulates not only the extent and progression of the biospecific lectin-induced cell-cell agglutination but also the topology of the contact seam. The IS dependence of contact separation in WGA-agglutinated cells is contrasted here with that reported for cells adhering in dextran solutions. The influence of IS change and pronase pre-treatment on contact pattern are consistent with predictions, from interfacial instability theory, of punctuate thinning of the aqueous layer separating bilayer membranes in close apposition.  相似文献   

2.
The equilibrium binding of ([125I]ceruloplasmin) ([125I]CP) to a specific receptor of human erythrocytes was investigated. It was shown that reaching the binding equilibrium is a slow process. A strong dependence of binding on Ca2+ concentration (from 0.1 to 1 mM) was revealed; the optimal values were achieved at millimolar concentrations of Ca2+.Mg2+ do not affect the binding of [125I]CP. Under conditions of optimal binding (0.01 M Tris-HCl buffer pH 7.4 containing 158 mM NaCl and 1 mM Ca2+, 4 degrees C), the values of constants for [125I]CP binding to intact erythrocytes (Kd = 1.0 nm) and to membrane fragments (Kd = 0.8 nM) as well as the number of binding sites (16.3 X 10(-15) mol per 40,000,000 erythrocytes) were determined. No ceruloplasmin transport across the erythrocyte membrane was observed. This finding and the similarity of Kd values for ceruloplasmin binding to membrane fragments and to intact erythrocytes indicate that the effect of ceruloplasmin on human erythrocytes is due to the protein molecule interaction with membrane receptors.  相似文献   

3.
The form of contact seam (whether a continuous parallel seam or membranes in spatially periodic contact) has been characterized for normal and for neuraminidase pretreated human erythrocytes following adhesion in solutions of polylysine in the molecular mass range 10-225 kDa at concentrations from 0.5 to 1.0 mg/mL. The adhesion contact seam was spatially periodic for all normal control cells in polylysine. The lateral separation of contacts decreased from 1.6 to 0.8 microns as the concentration of 225 kDa polylysine was increased threefold from the adhesion threshold value. The separation distance did not change further even at high polymer concentrations that increased the electrophoretic velocity to positive values over twice the modulus of the velocity of control cells. The probability of cell adhesion decreased at these high polymer concentrations. The lateral contact separation increased and cell adhesion decreased for cells pretreated with neuraminidase. Cell adhesion did not occur when neuraminidase reduced the cell electrophoretic velocity modulus by 30%. Following neuraminidase pretreatments that allowed a small amount of adhesion, the cell contact seam was continuous rather than spatially periodic. The results show that a procedure that increases (e.g., polymer concentration increase) or decreases (e.g., enzyme removal of polycation crosslinking site) attraction leads to shorter (to a limiting value) or longer lateral contact separation, respectively.  相似文献   

4.
The form of contact seam (whether a continuous parallel seam or membranes in spatially periodic contact) has been characterized for normal and for neuraminidase pretreated human erythrocytes following adhesion in solutions of polylysine in the molecular mass range 10–225 kDa at concentrations from 0.5 to 1.0 mg/mL. The adhesion contact seam was spatially periodic for all normal control cells in polylysine. The lateral separation of contacts decreased from 1.6 to 0.8 μm as the concentration of 225 kDa polylysine was increased threefold from the adhesion threshold value. The separation distance did not change further even at high polymer concentrations that increased the electrophoretic velocity to positive values over twice the modulus of the velocity of control cells. The probability of cell adhesion decreased at these high polymer concentrations. The lateral contact separation increased and cell adhesion decreased for cells pretreated with neuraminidase. Cell adhesion did not occur when neuraminidase reduced the cell electrophoretic velocity modulus by 30%. Following neuraminidase pretreatments that allowed a small amount of adhesion, the cell contact seam was continuous rather than spatially peridic. The results show that a procedure that increases (e.g., polymer concentration increase) or decreases (e.g., enzyme removal of polycation crosslinking site) attraction leads to shorter (to a limiting value) or longer lateral contact separation, respectively.  相似文献   

5.
Affinity-repulsion chromatography. Principle and application to lectins   总被引:3,自引:0,他引:3  
The interactions of proteins with their immobilized ligands in an electrically charged microenvironment were studied. The binding of lectins to erythrocytes and to affinity matrices was used as a model system. Lectins bind and agglutinate erythrocytes in the presence of at least 10 mM NaCl or 1 mM CaCl2, but not in deionized water. The salt dependence of the agglutination process is due to the ability of salts to provide counterions neutralizing the forces of repulsion between the electrostatic charges of similar sign present on the erythrocyte cell surface and on the lectins. The same salt dependence is observed for the binding of lectins to affinity matrices. These observations are the basis of a protein separation process coined affinity-repulsion chromatography in which the electrostatic charges present, or purposely introduced, on affinity matrices are exploited and allow the elution, by electrostatic repulsion, of proteins carrying electrostatic charges of the same sign as that of the matrix. In this process, proteins are loaded on the affinity matrix in a salt solution and eluted with deionized water. Affinity-repulsion chromatography has been successfully applied here to the isolation of several lectins. Its physicochemical basis, merits, and potential applications are discussed.  相似文献   

6.
The Escherichia coli UvrD protein (helicase II) is an SF1 superfamily helicase required for methyl-directed mismatch repair and nucleotide excision repair of DNA. We have characterized quantitatively the self-assembly equilibria of the UvrD protein as a function of [NaCl], [glycerol], and temperature (5-35 degrees C; pH 8.3) using analytical sedimentation velocity and equilibrium techniques, and find that UvrD self-associates into dimeric and tetrameric species over a range of solution conditions (t相似文献   

7.
The solution conformation of two peptides [1: PSGSNIISNLFKED; 2: GSSTLTALTTSVLKNNL] from human CD81 (hCD81) large extra-cellular loop (LEL) with known importance in the hepatitis C virus glycoprotein E2 (HCV-E2) binding interaction was characterized using circular dichroism spectroscopy. In addition, the solution structure of peptide 1 that contains a phenylalanine residue (F186 in hCD81) known to be critical in the binding interaction with HCV-E2 was determined using 1D and 2D 1H NMR spectroscopy. Both peptides are unstructured in water but begin forming significant helical conformation following the addition of 20% or more trifluoroethanol (v/v), a result consistent with their alpha-helical conformation found in the native protein. The CD data recorded as a function of pH and NaCl concentration are consistent with stabilization of the helical structure from electrostatic forces for both peptides. Peptide 1 is able to block the binding interaction of recombinant HCV-E2 (rHCV-E2) to hCD81 expressed on Molt-4 T cells at high concentrations (3.5 mM), a low affinity that we attributed to the random coil structure in water.  相似文献   

8.
A technique is described in which erythrocytes suspended in 1.1 ml of 145 mM NaCl, have been fused by electrofusion. The cells in suspension were brought into close contact by setting up a 3 MHz ultrasonic standing wave in a cylindrical cell container. The aluminium foil base of the container served both to transmit ultrasound and as an electrode for electrofusion. The electric pulse was generated by a capacitor discharge system. The electric field strength required to fuse cells increased as the ionic strength of the cell suspending phase increased. Cells in physiological saline fused at an electric field strength of 7.3 kV/cm with a 50 microseconds pulse.  相似文献   

9.
10.
The lysis of human erythrocytes by bile salts in buffer containing isotonic saline was dramatically enhanced by the addition of 5-10 mM calcium chloride. All bile acids tested showed this effect, with a marked increase in lysis occurring at 0.75 mM for deoxycholate, 1 mM for chenodeoxycholate, 2.5 mM for ursodeoxycholate and 5.5 mM with cholate in the presence of 10 mM calcium chloride. The effect appeared to be specific for calcium; strontium chloride and magnesium chloride gave no stimulatory effect. The increased lysis of the erythrocytes in the presence of 1 mM deoxycholate and 1-10 mM calcium chloride was not associated with increased uptake of the bile salt by the cells (measured with [14C]deoxycholate). Using erythrocytes previously labelled with [3H]cholesterol, there was no evidence of an enhanced removal of that membrane component in the presence of calcium and deoxycholate, compared to deoxycholate alone. The sensitivity of the cells to the effect of calcium in the presence of 1 mM deoxycholate increased with the length of time of their storage at 4 degrees C. The sensitivity returned to that of fresh cells after incubation at 37 degrees C with 30 mM adenosine plus 25 mM glucose, but this treatment did not further diminish the lysis. Lysis in the presence of 10 mM calcium chloride and 1 mM deoxycholate was partially blocked by increasing the KCl concentration at the expense of NaCl. The maximum effect occurred with a buffer comprising 100 mM KCl/50 mM NaCl. A more dramatic reduction in the lysis followed the incorporation of the calcium chelator, quin2, into the cells. The lysis induced by 1 mM deoxycholate in the presence of calcium was reduced by 80% in quin-2-loaded cells compared to controls. The data suggest that bile acids can promote the influx of calcium into erythrocytes, leading to lysis as a result of the efflux of intracellular potassium and/or the uptake of sodium from the incubation medium. The data further suggest that cellular effects may occur at lower bile acid concentrations than that thought to be required for detergent damage.  相似文献   

11.
Hsp70 and a 54 kDa osmotic stress protein (osp54) were induced in isolated tissues of anadromous Atlantic salmon (Salmo salar) upon exposure to hyperosmotic conditions. Incubation of branchial lamellae, hepatic tissue, and erythrocytes in medium supplemented with 200-600 mM NaCl dramatically reduced protein synthesis. Although general protein synthesis remained depressed following transfer of tissues from 450 mM supplemental NaCl to iso-osmotic medium, hsp70 was prominently induced in branchial lamellae and hepatic tissue. Accumulation of hsp70 mRNA and a decrease in actin mRNA suggest preferential upregulation of the hsp70 gene. Induction of osp54 was observed in branchial lamellae and erythrocytes, but not in hepatic tissue, during exposure to 75-125 mM supplemental NaCl. Use of glycerol in place of NaCl to create hyperosmotic conditions stimulated induction of hsp70 in branchial lamellae. Substitution with mannitol resulted in induction of osp54 in both branchial lamellae and erythrocytes. The solute-specific and temporal patterns of response suggest that hsp70 and osp54 might function in concert to restore osmotic homeostasis and renature proteins destabilized or denatured during the early stages of osmotic shock.  相似文献   

12.
Zu Y  Di Bernardo S  Yagi T  Hirst J 《Biochemistry》2002,41(31):10056-10069
The redox properties of the [2Fe-2S] cluster in the 24 kDa subunit of bovine heart mitochondrial NADH:ubiquinone oxidoreductase (complex I) and three of its homologues have been defined using protein-film voltammetry. The clusters in all four examples display characteristic, pH-dependent redox transitions, which, unusually, can be masked by high ionic strength conditions. At low ionic strength (10 mM NaCl) the reduction potential varies by approximately 100 mV between high and low pH limits (pH 5 and 9); thus the redox process is not strongly coupled and is unlikely to form part of the mechanism of energy transduction in complex I. The pH dependence was shown to result from pH-linked changes in protein charge, due to nonspecific protonation events, rather than from the coupling of a specific ionizable residue, and the ionic strength dependence at high and low pH was modeled using extended Debye-Hückel theory. The low potential of the 24 kDa subunit [2Fe-2S] cluster, out of line with the potentials of the other iron-sulfur clusters in complex I, is suggested to play a role in coupling reducing equivalents at the catalytic active site. Finally, the validity of using the [2Fe-2S] cluster in an isolated subunit, as a mechanistic basis for coupled proton-electron transfer in intact complex I, is evaluated.  相似文献   

13.
The binding of ciprofloxacin to natural and synthetic polymeric DNAs was investigated at different solvent conditions using a combination of spectroscopic and hydrodynamic techniques. In 10 mM cacodylate buffer (pH 7.0) containing 108.6 mM Na(+), no sequence preferences in the interaction of ciprofloxacin with DNA was detected, while in 2 mM cacodylate buffer (pH 7.0) containing only 1.7 mM Na(+), a significant binding of ciprofloxacin to natural and synthetic linear double-stranded DNA was observed. At low ionic strength of solution, ciprofloxacin binding to DNA duplex containing alternating AT base pairs is accompanied by the largest enhancement in thermal stability (e.g. DeltaT(m) approximately 10 degrees C for poly[d(AT)].poly[d(AT)]), and the most pronounced red shift in the position of the maximum of the fluorescence emission spectrum (lambda(max)). Similar red shift in the position of lambda(max) is also observed for ciprofloxacin binding to dodecameric duplex containing five successive alternating AT base pairs in the row. On the other hand, ciprofloxacin binding to poly[d(GC)].poly[d(GC)], calf thymus DNA and dodecameric duplex containing a mixed sequence is accompanied by the largest fluorescence intensity quenching. Addition of NaCl does not completely displace ciprofloxacin bound to DNA, indicating the binding is not entirely electrostatic in origin. The intrinsic viscosity data suggest some degree of ciprofloxacin intercalation into duplex.  相似文献   

14.
Marked differences were found among 28 finger millet genotypes(Eleusine coracana Gaertn.) in acquired tolerance to osmoticstress as assessed by the recovery of root growth from severestress [-1·2 MPa polyethylene glycol, (PEG) or 400 mMNaCl]. However, these differences in tolerance were observedonly when the seedlings were subjected to a preceding mild inductionstress (-0·6 MPa PEG or 200 mM NaCl). In two contrastinggenotypes, synthesis of stress-induced proteins was studied.Proteins with apparent molecular weight of 70-72, 52, 37, 34and 23 kDa were synthesized in the highly responsive genotype(GE 415) and poorly responsive (VL 481) genotype following amild induction stress (200 mM NaCl). However, GE-415 synthesizeda 54 kDa protein that was not observed in VL-481. Addition ofabscisic acid (ABA) to the induction medium containing 200 mMNaCl enhanced the acquired tolerance of finger millet seedlingsover those without ABA in association with the appearance ofseveral ABA-responsive proteins. GE-415 required much less ABAthan VL-481 to obtain the same response. With 10 µM ABA+ 200 mM, A NaCl induction stress, GE-415 had significantlyhigher endogenous ABA. In association with higher levels ofABA, GE-415 had greater recovery root growth following severestress from 600 mM NaCl. Pretreatment with 10 µM ABA +200 mM NaCl induced several proteins with apparent molecularweights of 70-72, 54, 45, 36, 29 and 21 kDa in both genotypes.Qualitatively, GE-415 synthesized a unique 23-24 kDa proteinand quantitatively there was significantly more of the 21 kDaprotein in GE-415 compared to VL-481. The results indicate thatthe synthesis of stress proteins is correlated with the observedvariation in acquired tolerance of the two genotypes.Copyright1995, 1999 Academic Press Eleusine coracana Gaertn., salinity, polyethylene glycol, stress proteins, ABA, ABA-responsive proteins, finger millet seedlings  相似文献   

15.
We investigated effects of low Mg and moderately raised NaCl, as occurs in plantations irrigated with tertiary municipal effluent in New Zealand, on growth, Mg nutrition and photosynthetic activity of Pinus radiata D. Don seedlings grown in nutrient solutions with a Perlite medium. Seedlings were grown with either sufficient (0.35 mM; H[Mg]) or limited (0.033 mM; L[Mg]) Mg supply, without NaCl or with NaCl addition (8.7 mM; +[NaCl]). After 30 weeks, seedlings grown at L[Mg] displayed severe Mg deficiency symptoms, and had significantly less biomass than those at H[Mg]. While NaCl addition had an adverse effect on seedling growth at H[Mg], it increased growth at L[Mg]. The +[NaCl] treatment greatly increased the Mg uptake rates, which were associated with increased stomatal conductance and increased root to shoot ratio. Magnesium deficiency reduced the rates of light-saturated photosynthesis and stomatal conductance, but not the quantum efficiency of photosystem II, which was reduced mainly by the +[NaCl] treatment, especially at H[Mg]. Our study clearly indicated that NaCl addition could counteract the impact of low Mg supply by enhancing Mg uptake from the rooting medium.  相似文献   

16.
Equilibrium binding studies and viscosity experiments are described that characterize the interaction of delta- and lambda-[Ru(o-phen)3]2+ with calf thymus DNA. The mode of binding of these compounds to DNA is a matter of controversy. Both isomers of [Ru(o-phen)3]2+ were found to bind but weakly to DNA, with binding constants of 4.9 (+/- 0.3) x 10(4) M-1 and 2.8 (+/- 0.2) x 10(4) M-1 determined for the delta and lambda isomers, respectively, at 20 degrees C in a solution containing 5 mM Tris-HCl (pH 7.1) and 10 mM NaCl. We determined that the quantity delta log K/delta log [Na+] equals 1.37 and 1.24 for the delta and lambda isomers, respectively. Application of polyelectrolyte theory allows us to use these values to show quantitatively that both the delta and lambda isomers are essentially electrostatically bound to DNA. Viscosity experiments show that binding the lambda isomer does not alter the relative viscosity of DNA to any appreciable extent, while binding of the delta isomer decreases the relative viscosity of DNA. From these viscosity results, we conclude that neither isomer of [Ru(o-phen)3]2+ binds to DNA by classical intercalation.  相似文献   

17.
It was recently proposed that alpha-D-glucose 6-phosphate may undergo enzyme-to-enzyme channelling between glucokinase and phosphoglucoisomerase in rat pancreatic islets. The present study aims at exploring whether a different situation prevails in cells deprived of glucokinase, namely in erythrocytes. At anomeric equilibrium, the ratio between D-[2-3H]glucose and D-[5-3H]glucose conversion to 3HOH was lower in rat erythrocytes incubated for 60 min at 4 degrees C in the presence of 2.8 mM, rather than 8.3 mM, D-glucose. This coincided with both a greater relative increase in beta-D-[5-3H]glucose, as compared to alpha-D-[5-3H]glucose, conversion to 3HOH and an increase in the beta/alpha ratio for 3HOH generation from D-[5-3H]glucose in response to an increase in the anomeric concentration from 2.8 to 8.3 mM, the suppression of the difference between the beta/alpha ratios for 3HOH generation from D-[2-3H]glucose and D-[5-3H]glucose in the erythrocytes incubated at 8.3 mM, as distinct from 2.8 mM, alpha- and beta-D-glucose, and a [2-3H]/[5-3H] ratio for 3HOH generation lower than unity in erythrocytes exposed to alpha-D-glucose but not significantly different from unity in the presence of beta-D-glucose. These findings emphasize the relevance of alpha-D-glucose 6-phosphate channelling between hexokinase and phosphoglucoisomerase as a determinant of the difference between D-[2-3H]glucose and D-[5-3H]glucose conversion to 3HOH, and reveal that the regulation of such a tunnelling process by the concentration of the D-glucose represents, in rat erythrocytes, a mirror image of that observed in rat pancreatic islets. The regulation of this process thus tightly depends on the identity of the hexokinase enzyme mainly responsible for the phosphorylation of D-glucose in distinct cell types.  相似文献   

18.
Proton conductivities in bulk solution (lambda(H)) and single-channel proton conductances (g(H)) in two different stereoisomers of the dioxolane-linked gramicidin A channel (the SS and RR dimers) were measured in a wide range of bulk proton concentrations ([H], 0.1-8000 mM). Proton mobilities (micro(H)) in water as well as in the SS and RR dimers were calculated from the conductivity data. In the concentration range of 0.1-2000 mM, a straight line with a slope of 0.75 describes the log (g(H))-log ([H]) relationship in the SS dimer. At [H] > 2000 mM, saturation is followed by a decline in g(H). The g(H)-[H] relationship in the SS dimer is qualitatively similar to the [H] dependence of lambda(H). However, the slope of the straight line in the log(lambda(H))-log([H]) plot is 0.96, indicating that the rate-limiting step for proton conduction through the SS dimer is not the diffusion of protons in bulk solution. The significant difference between the slopes of those linear relationships accounts for the faster decline of micro(H) as a function of [H] in the SS dimer in relation to bulk solution. In the high range of [H], saturation and decline of g(H) in the SS dimer can be accounted for by the significant decrease of micro(H) in bulk solution. At any given [H], g(H) in the RR dimer is significantly smaller than in the SS. Moreover, the g(H)-[H] relationship in the RR stereoisomer is qualitatively different from that in the SS. Between 1 and 50 mM [H], g(H) can be fitted with an adsorption isotherm, suggesting the presence of a proton-binding site inside the pore (pK(a) approximately 2), which limits proton exit from the channel. At 100 mM < [H] < 3000 mM, g(H) increases linearly with [H]. The distinctive shape of the g(H)-[H] relationship in the RR dimer suggests that the channel can be occupied simultaneously by more than one proton. At higher [H], the saturation and decline of g(H) in the RR dimer reflect the properties of micro(H) in bulk solution. In the entire range of [H], protons seem to cross the SS and RR channels via a Grotthuss-like mechanism. The rate-limiting step for proton transfer in the SS dimer is probably the membrane-channel/bulk solution interface. It is also proposed that the smaller g(H) in the RR dimer is the consequence of a different organization and dynamics of the H-bonded network of water molecules inside the pore of the channel, resulting in a slower proton transfer and multiple pore occupancy by protons.  相似文献   

19.
The pathway by which L-lactate (Lac) crosses the plasma membrane of isolated human neutrophils was investigated. The influx of [14C]Lac from a 2 mM Lac, 145 mM Cl-, 5.6 mM glucose medium was approximately 1.5 meq/liter of cell water.min and was sensitive to the organomercurial agent mersalyl (apparent Ki approximately 20 microM), to alpha-cyano-4-hydroxycinnamate (CHC), the classical inhibitor of monocarboxylate transport in mitochondria, and to UK-5099 (apparent Ki approximately 40 microM), a more potent analogue of CHC. Transport was also strongly blocked (greater than 80%) by 1 mM of either 3,5-diiodosalicylic acid, MK-473 (an indanyloxyacetate derivative), or diphenyl-amine-2-carboxylate, and by 0.4 mM pentachlorophenol, but not by 1 mM ethacrynic acid, furosemide, or the disulfonic stilbenes SITS or H2DIDS. One-way [14C]Lac efflux from steady-state cells amounted to approximately 6 meq/liter.min and was likewise affected by the agents listed above. Influx, which was membrane potential insensitive and Na+ independent, displayed a strong pH dependence: extracellular acidification enhanced uptake while alkalinization inhibited the process (pK' approximately 5.7 at 2 mM external Lac). The rate of [14C]Lac influx was a saturable function of external Lac, the Km being approximately 7 mM. Steady-state cells exhibited an intracellular Lac content of approximately 5 mM and secreted lactic acid into the bathing medium a a rate of approximately 4 meq/liter.min. Secretion was completely suppressed by 1 mM mersalyl which inactivates the carrier, leading to an internal accumulation of Lac. That the Lac carrier truly mediates an H+ + Lac- cotransport (or formally equivalent Lac-/OH- exchange) was documented by pH-stat techniques wherein an alkalinization of poorly buffered medium could be detected upon the addition of Lac; these pH changes were sensitive to mersalyl. Thus, the Lac carrier of neutrophils possesses several features in common with other monocarboxylate transport systems in erythrocytes and epithelia.  相似文献   

20.
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   

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