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 46 毫秒
1.
Simulations for the moving-boundary ultracentrifuge behavior of Helix pomatia alpha-hemocyanin have been developed, based on modification of published models. In the present treatment, it has been assumed that the protein system is extensively microheterogeneous, with respect to its whole-half molecule reactivity. It has also been assumed that all such association and dissociation reactions can proceed to equilibrium in a time appreciably shorter than that which would be required to separate nonreacting half molecules from nonreacting whole molecules. Predictions based on this model agree well with reported experimental findings of nonequilibration of fractionated material, apparent independence of whole-to-half molecule concentration ratios on total concentration in sedimentation experiments, and stopped-flow dilution reactivity over a wide range of sedimentation coefficients.  相似文献   

2.
Two independent relaxation kinetics methods were used to study samples of α-hemocyanin kindly furnished to us by members of the Biochemical Laboratory of the University of Groningen. A Durrum-Gibson stopped-flow apparatus was used to obtain concentration-jump data in the light-scattering mode. A recently developed pressurejump light-scattering apparatus was used to obtain completely independent data. The studies were made in 0.1 m acetate buffer at pH 5.7 containing 0.4 m NaCl, conditions under which equilibrium light-scattering studies had been reported by Engelborghs and Lontie (1973, J. Mol. Biol., 77, 577–587). In the companion paper (Kegeles, 1977, Arch. Biochem. Biophys., 180, 530–536), a model is proposed, consisting of a system containing a mixture of reactive and unreactive whole molecules, from which data are derived for the formation constant of whole molecules from halves and the fraction of material which is capable of undergoing reaction. The present study uses this estimate of this fraction of reactive material to permit the evaluation of overall rate constants and equilibrium constants. When the estimate of 65% of reactive material derived without making nonideality corrections is applied to the kinetics data, very satisfactory agreement is obtained between the equilibrium constant acquired from equilibrium data and the equilibrium constants derived from the kinetics data.  相似文献   

3.
The role of charged residues on the surface of plastocyanin from the cyanobacterium Phormidium laminosum in the reaction with soluble cytochrome f in vitro was studied using site-directed mutagenesis. The charge on each of five residues on the eastern face of plastocyanin was neutralized and/or inverted, and the effect of the mutation on midpoint potentials was determined. The dependence of the overall rate constant of reaction, k(2), on ionic strength was investigated using stopped-flow spectrophotometry. Removing negative charges (D44A or D45A) accelerated the reaction and increased the dependence on ionic strength, whereas removing positive charges slowed it down. Two mutations (K46A, K53A) each almost completely abolished any influence of ionic strength on k(2), and three mutations (R93A, R93Q, R93E) each converted electrostatic attraction into repulsion. At low ionic strength, wild type and all mutants showed an inhibition which might be due to changes in the interaction radius as a consequence of ionic strength dependence of the Debye length or to effects on the rate constant of electron transfer, k(et). The study shows that the electrostatics of the interaction between plastocyanin and cytochrome f of P. laminosum in vitro are not optimized for k(2). Whereas electrostatics are the major contributor to k(2) in plants [Kannt, A., et al. (1996) Biochim. Biophys. Acta 1277, 115-126], this role is taken by nonpolar interactions in the cyanobacterium, leading to a remarkably high rate at infinite ionic strength (3.2 x 10(7) M(-1) s(-1)).  相似文献   

4.
Gizzard heavy meromyosin (HMM) sediments in the ultracentrifuge as a single peak, whose sedimentation coefficient (S20,w) decreases from 9 to 7.5 S upon increasing the NaCl concentration from 0.02 to 0.3 M. This decrease is accompanied by a parallel increase in Mg2+-ATPase activity, suggesting that both changes have a common molecular basis. Phosphorylation decreases S20,w and increases ATPase activity, while ATP increases S20,w. Sedimentation equilibrium studies indicate that HMM undergoes no detectable aggregation at 0.02 or 0.4 M NaCl, remaining monomeric with a molecular weight of 3.4 X 10(5). In contrast, S20,w of subfragment 1 does not change with changes in ionic strength, and its ATPase activity does not decrease at low ionic strengths. Electron micrographs of samples of HMM prepared at low ionic strength show that up to half of the molecules are flexed, i.e. the heads are bent at the neck and project back toward the tail, while the remaining molecules have either one or both of the heads pointing away from the tail. In samples prepared at high ionic strength only about 10% of the molecules are flexed. There is a linear relationship between the fraction of flexed molecules and S20,w, with no significant bending or folding of the tail and no detectable change in the shape of the heads. This correlation suggests that the changes in ATPase activity and S20,w may be a result of the reorientation of the heads.  相似文献   

5.
6.
Nucleosome dissociation at physiological ionic strengths.   总被引:5,自引:2,他引:3       下载免费PDF全文
Monomer nucleosomes purified on isokinetic sucrose gradients are shown to dissociate into component DNA and histones at physiological ionic strength upon dilution to a DNA concentration below 20 microgram/ml. The starting material is 11S, contains 145-190 BP DNA, and equimolar amounts of the four core histones with slightly less H1. Dilution of monomers in the presence of 0.14 M NaCl results in the rapid conversion of 10-40% of the 3H thymidine labeled material from 11S to 5S (5S is coincident with the S value of monomer length DNA). The proportion of nucleosomes which dissociate increases with increasing NaCl concentration between 0.15 M and 0.35 M and decreases with increasing DNA concentration above 1 microgram/ml. Recycling 11S monomers, which remain after dissociation, through a second dilution in salt generates an equivalent proportion of 5S material as seen after the initial dilution. Thus, the dissociation does not result from special properties of a subset of nucleosomes. An equilibrium between intact monomer and free DNA and histones appears to be rapidly established under the conditions described and the dissociated DNA will reassociate with histones to form 11S monomers if conditions of high DNA concentration and low ionic strength are established.  相似文献   

7.
8.
It was shown previously that a significant fraction of the myosin crossbridges is attached to actin in the skinned rabbit psoas fibers under relaxed conditions at low ionic strength and low temperature (Brenner, B., M. Schoenberg, J. M. Chalovich, L. E. Greene, and E. Eisenberg. 1982. Proc. Natl. Acad. Sci. USA. 79:7288-7291; Brenner, B., L. C. Lu, and R. J. Podolsky. 1984. Biophys. J. 46:299-306). In the present work, the structure of the attached crossbridges in the relaxed state between ionic strengths of 20 and 100 mM, as compared with that in the rigor state, is further examined by equatorial x-ray diffraction. Mass distributions projected along the fiber axis are reconstructed based on the first five equatorial reflections such that the spatial resolution is 128 A. The fraction of crossbridges attached under relaxed conditions are estimated to be in the range of 30% (at 100 mM ionic strength) and 60% (at 20 mM). The reconstructed density maps suggest that in the relaxed state, upon attachment the part of the crossbridge that centers around the thin filament is small, and the attachment does not significantly alter the center of mass of the myosin head distribution around the thick filament backbone. In contrast, accretion of mass in the rigor state occurs in a wider region surrounding the thin filament. In this case, mass in the surface region of the thick filament backbone is shifted slightly outward, probably by approximately 10 A. A schematic model for interpreting the present data is presented.  相似文献   

9.
Helix pomatia α-hemocyanin can dissociate stepwise into 12-size, 110-size and 120-size molecules. Both 110-size and 120-size molecules can occur in two isomeric forms, differing considerably in sedimentation behaviour.The effects of pH, ionic strength, temperature, Ca2+ concentration and oxygen pressure on these dissociation and isomerization steps were investigated systematically by sedimentation analysis.Dissociation and isomerization are favoured by increasing pH or temperature. Changes in ionic strength affect each step differently. Calcium ions are extremely effective in preventing dissociation and isomerization at low ionic strength, but this stabilizing ability diminishes at higher ionic strengths. Oxygen binding shifts the pH-dependent dissociation of whole into 12-size molecules to higher pH. Oxygen has no effect on the other dissociation steps. Intermolecular interactions appear to be predominantly electrostatic.  相似文献   

10.
Kinesin undergoes a 9 S to 6 S conformational transition.   总被引:6,自引:0,他引:6  
Addition of NaCl or KCl in the presence of 50 nM ATP induces a shift in the sedimentation coefficient (apparent S20,w) of kinesin from 9.4 S at low ionic strength to 6.5 S at high ionic strength. The midpoint for the transition occurs at ionic strength values of 0.39, 0.25, and 0.18 for pH values of 6.3, 6.9, and 8.3, respectively. Gel filtration experiments indicate that the transition to the 6.5 S species is accompanied by a decrease in the diffusion coefficient. Under all conditions which were tested, the 64-kDa beta subunits comigrate with the 120-kDa alpha subunits without any evidence for dissociation of the alpha 2 beta 2 complex. These results are consistent with the change in sedimentation coefficient being due to a conformational transition between a folded form at low ionic strength and an extended form at high ionic strength. This conformational transition is not significantly affected by the nature of the nucleotide bound at the active site since similar results are obtained both in the presence of excess EDTA, which removes the bound ADP, and after replacement of the bound ADP with adenosine 5'-(beta,gamma-imino)triphosphate. The alpha 2 form of kinesin, which lacks the beta subunits, undergoes a similar transition between a 6.7 S form at low ionic strength and a 5.1 S form at high ionic strength with a midpoint for the transition at an ionic strength of 0.5 at pH 6.9. Electron microscopic observation also indicates a transition between a folded conformation at low ionic strength and an extended conformation at high ionic strength for both the alpha 2 beta 2 and alpha 2 species.  相似文献   

11.
Damaged starch characterisation by ultracentrifugation   总被引:1,自引:0,他引:1  
The relative molecular size distributions of a selection of starches (waxy maize, pea and maize) that had received differing amounts of damage from ball milling (as quantified by susceptibility to alpha-amylase) were compared using analytical ultracentrifugation. Starch samples were solubilised in 90% dimethyl sulfoxide, and relative size distributions were determined in terms of the apparent distribution of sedimentation coefficients g*(s) versus s(20,w). For comparison purposes, the sedimentation coefficients were normalised to standard conditions of density and viscosity of water at 20 degrees C, and measurements were made with a standard starch loading concentration of 8 mg/mL. The modal molecular size of the native unmilled alpha-glucans were found to be approximately 50S, 51S and 79S for the waxy maize, pea and maize amylopectin molecules, respectively, whilst the pea and maize amylose modal molecular sizes were approximately 14S and approximately 12S, respectively. As the amount of damaged starch increased, the amylopectin molecules were eventually fragmented, and several components appeared, with the smallest fractions approaching the sedimentation coefficient values of amylose. For the waxy maize starch, the 50S material (amylopectin) was gradually converted to 14S, and the degradation process included the appearance of 24S material. For the pea starch, the situation was more complicated than the waxy maize due to the presence of amylose. As the amylopectin molecules (51S) were depolymerised by damage within this starch, low-molecular-weight fragments added to the proportion of the amylose fraction (14S)--although tending towards the high-molecular-weight region of this fraction. As normal maize starch was progressively damaged, a greater number of fragments appeared to be generated compared to the other two starches. Here, the 79S amylopectin peak (native starch) was gradually converted into 61 and 46S material and eventually to 11S material with a molecular size comparable to amylose. Amylose did not appear to be degraded, implying that all the damage was focused on the amylopectin fraction in all three cases. Specific differences in the damage profiles for the pea and maize starches may reflect the effect of lipid-complexed amylose in the maize starch.  相似文献   

12.
1. The conformation of bovine microtubule protein prepared by cycles of assembly and disassembly in the presence of glycerol has been studied by near-u.v. circular dichroism (c.d.) over a range of protein concentrations. The effects on the conformational properties of ionic strength and of a pH range from 6 to 7.5 have been correlated with the known oligomeric composition of microtubule protein preparations, as determined by the sedimentation behaviour of this preparation [Bayley, Charlwood, Clark & Martin (1982) Eur. J. Biochem. 121, 579–585]. 2. The formation of 30S oligomeric ring species, either by decreasing ionic strength at pH6.5 or by changing pH in the presence of 0.1m-NaCl, correlates with a significant change in tubulin c.d. Formation of 18S oligomer by changing pH at ionic strength 0.2 produced no comparable effect. The c.d. of tubulin dimer itself is not affected by ionic strength and pH over the same range. 3. The results are interpreted as a small conformational adjustment between tubulin and specific microtubule-associated proteins on forming 30S oligomeric species, due to interaction with the high-molecular-weight-group proteins. The possible significance of this is discussed with respect to microtubule assembly in vitro. 4. By using this conformational parameter, together with equilibrium and kinetic light-scattering studies, the sensitivity of glycerol-cycled microtubule protein to dilution is shown to be strongly pH-dependent, the oligomers being much more stable at pH6.4 than at pH6.9. 5. Oligomeric complexes of tubulin with microtubule-associated proteins show marked stability under conditions similar to those for efficient microtubule assembly in vitro. Oligomeric material therefore must be incorporated directly during assembly in vitro from microtubule protein.  相似文献   

13.
Neurofilament proteins purified from bovine spinal cord were characterized by sedimentation studies in aqueous buffers. In 10 mM Tris, pH 8, the middle molecular weight neurofilament protein (NF-M) has a sedimentation coefficient, S20,w, of 2.6 S. Sedimentation equilibrium data shows considerable nonideality; extrapolation to infinite dilution and correction for the primary charge effect yield a molecular weight of 1.09 X 10(5), indicative of a monomeric structure. When the ionic strength was increased, the sedimentation coefficient increased slightly, and the protein began to form larger aggregates. Reconstitution of short intermediate filaments was observed upon dialysis of denatured NF-M versus a reconstitution buffer. A circular dichroism spectrum of NF-M in 10 mM Tris was typical of alpha + beta proteins. High molecular weight neurofilament protein (NF-H) showed a considerable tendency to aggregate in 10 mM Tris, but a principal species with a sedimentation coefficient of 3.2 S was observed, and sedimentation equilibrium data also suggest a monomeric structure.  相似文献   

14.
A DNA polymerase-endogenous template complex was isolated from nuclear heads of bull spermatozoa. The buoyant density of the complex was 1.15 g/cm 3. The sedimentation coefficient of the nuclear DNA polymerase isolated from the complex was higher at low ionic strength, but approached 3.4S when centrifuged in a medium containing 2M-KCl. Activated exogenous DNA increased polymerase activity. Only very low activities were detected with synthetic templates such as poly(A).(dT)12-18 and poly(dT).poly(A). The nuclear reaction was stimulated by 150mM-KCl and was slightly inhibited by N-ethylmaleimide; it was resistant to actinomycin D, netropsin and ethidium bromide. Another DNA polymerase, highly sensitive to ethidium bromide, was extracted from the mitochondira-rich middle-piece fraction. Its sedimentation coefficient was close to 9S, but fell to approx. 4S in high-ionic-strength medium.  相似文献   

15.
The self-association of alfalfa mosaic virus coat protein was studied by sedimentation analysis and electron microscopy under a wide range of conditions. In the depolymerized state the protein exists as a molecular species with a sedimentation constant of roughly 3 S and with a molecular weight of (48.4 ± 1.1) × 103. This value is, within experimental error, twice the value of the monomer (van Beynum, 1975). The dimer has a very stable configuration, as no evidence was found for a monomer-dimer equilibrium between pH values of 3 and 9 and values of ionic strength up to 1.0. One main type of association product (30 S) was found with a molecular weight of (1.48 ± 0.03) × 106. Therefore this particle accomodates 30 dimers which are arranged according to a point group symmetry of 532. The orientation of the 30 dimers within the icosahedral lattice must be such that lattice dyads coincide with the 2-fold axes of the dimers. Micrographs of the 30 S particles show a diameter of about 123 Å; analysis of linear arrays of these particles shows that at low resolution the particle is a hollow sphere with an average coat thickness of about 40 Å.The influence of pH, ionic strength, protein concentration and the type of buffer on the polymerization was determined to some extent and is discussed. The assembly of dimers into the icosahedral particle is an entropy-driven process (Lauffer, 1975); this is concluded from studying the temperature-dependence of the free energy change. Under favourable conditions (phosphate buffer pH 5.5 and ionic strength 0.5) the average enthalpy and entropy changes for the insertion of one dimer into the lattice are about 6.4 kilocalories per mole and 50 entropy units, respectively, based on the unit mole fraction.  相似文献   

16.
The effect of 2,3-diphospho-D-glycerate on the sedimentation coefficient of carbon monoxide hemoglobin was correlated with the fraction of rapidly reacting hemoglobin observed subsequent to flash photolysis at 23 degrees C at pH 7.30 in buffers of 0.1 M ionic strength. Concentrations of the organic phosphate up to about 5 mM resulted in an increase in S20,w, consistent with an increase in the fraction of tetrameric hemoglobin. A decrease in rapidly reacting hemoglobin parallelled the increase in the sedimentation coefficient. Between 5 and 20 mM 2,3-diphosphoglycerate, S20,w decreased, suggesting that dissociation to dimers was enhanced. An increase in rapidly reacting hemoglobin was also observed in this concentration range. Similar sedimentation results were obtained with oxyhemoglobin at pH 7.00 and carbon monoxide hemoglobin at pH 7.06. Assuming single binding sites on each species, the dissociation constants for 2,3-diphosphoglycerate binding to tetrameric and dimeric HbCO are 0.2-0.3 mM and 2-5 mM at pH 7.30. This biphasic effect of this physiologically important organic phosphate on the state of aggregation of R state hemoglobin has not been previously reported, but it is similar to that previously noted with inositol hexaphosphate, which enhanced tetramer formation at low concentrations, while at higher concentrations it promoted hemoglobin dissociation to dimers (White, S. L. (1976) J. Biol. Chem. 251, 4763-4769; Gray, R. D. (1980) J. Biol. Chem. 255, 1812-1818).  相似文献   

17.
The formation of higher order structures by nucleosome oligomers of graded sizes with increasing ionic strength has been studied in solution, by measuring sedimentation coefficients. Nucleosome monomers and dimers show no effect of ionic strength at the concentrations used, while trimers to pentamers show a linear dependence of the logarithm of sedimentation coefficient upon the logarithm of ionic strength between 5 and 25 mm, but no dependence above 25 mm. Between pentamer and hexamer a change occurs and the linear relationship is observed up to ionic strength 125 mm with hexamer and above.The simple power-law dependence of the sedimentation coefficient upon the ionic strength (sIn) is observed up to nucleosome 30mers, but by 60mer a jump in the sedimentation coefficient occurs between ionic strengths 45 and 55 mm, with the power-law applying both above and below the jump. Removal of histone H1 and non-histone proteins lowers the overall sedimentation rate and abolishes the jump.Cross-linking large oligomers at ionic strength 65 mm stabilizes the structure in the conformation found above the jump, leading to a simple power-law dependence throughout the range of ionic strength for cross-linked material. Cleavage of the cross-links restores the jump, presumably by allowing the conformational transition that causes it. Large oligomers are indistinguishable in sedimentation behaviour whether extracted from nuclei at low ionic strength or at 65 mm and maintained in the presence of salt.We interpret these results, together with the detailed electron microscopic studies reported by Thoma et al. (1979) under similar salt conditions, as showing the histone H1-dependent formation of superstructures of nucleosomes in solution induced by increasing ionic strength. The unit of higher order structure probably contains five or six nucleosomes, leading to the change in stability with hexamer. Although this size corresponds to the lower limit of size suggested for “superbeads” (Renz et al., 1977), we see no evidence that multiples of six nucleosomes have any special significance as might be predicted if superbeads had any structural importance. Rather, our results are compatible with a continuous pattern of condensation, such as a helix of nucleosomes (see e.g. Finch & Klug, 1976). The jump in sedimentation observed between ionic strengths 45 and 55 mm, together with the effect of cross-linking, suggests the co-operative stabilization of this structure at higher ionic strengths. A plausible hypothesis is that the turns of the solenoid are not tightly bonded in the axial direction below 45 mm, but come apart due to the hydrodynamic shearing forces in the larger particles leading to less compact structures with slower sedimentation rates. Above 55 mm the axial bonding is strong enough to give a stable structure of dimensions compatible with the 30 nm structures observed in the cell nucleus.  相似文献   

18.
Mitochondrial DNA polymerase (DNA polymerase mt) exists in two active forms. DNA polymerase present in crude extract (M-I) and ammonium sulfate precipitate (M-II) stages of purification sediments at 12.1S. The enzyme at the M-II stage of purification has a molecular weight of approximately 250,000 as determined by Sephadex G-200 chromatography in buffers of low ionic strength. In buffers containing 0.15 m NaCl, the enzyme sediments at 9.4S and has a molecular weight of approximately 190,000. When the enzyme is further purified on diethylaminoethyl cellulose (M-III stage of purification), the 9.4S activity predominates. Addition of a polymerase-free fraction from the M-III stage of purification changes the sedimentation coefficient of the enzyme from 9.4 to 12.1S.  相似文献   

19.
1. A study was made of the sedimentation properties of purified preparations of the rapidly labelled RNA in the nucleus and the cytoplasm of the HeLa cell. The sedimentation of the rapidly labelled nuclear RNA was very sensitive to changes in ionic strength and bivalent cation concentration. Under the conditions usually used in sucrose-density-gradient centrifugation the rapidly labelled nuclear RNA showed extreme polydispersity, and much of it sedimented more rapidly than the 28s RNA. At low ionic strength and after removal of Mg(2+), however, the rapidly labelled nuclear RNA sedimented as a single peak at about 16s. The conversion of the polydisperse material into the 16s form did not involve degradation of the RNA, since the effect could be reversed by increasing the ionic strength of the solution. 2. The cytoplasm did not contain any RNA that showed polydisperse sedimentation under the usual conditions of sucrose-density-gradient centrifugation, or that had the same sensitivity as the rapidly labelled nuclear RNA to changes in ionic strength. All the radioactivity in the cytoplasmic RNA sedimented with the 28s, 16s and 4s components over a wide range of physical conditions, but these components did contain a labelled fraction with some of the features of the rapidly labelled nuclear RNA on columns of methylated albumin on kieselguhr. 3. In both nucleus and cytoplasm the RNA detected by ultraviolet absorption could also be converted into a 16s form by removal of bivalent cations at low ionic strength; this effect was again, within certain limits, reversible. The nuclear RNA as a whole was more susceptible to changes in ionic strength than the cytoplasmic RNA. 4. It thus appears that all the RNA in the cell, except the 4s RNA, can be prepared, without degradation, as a single peak sedimenting at about 16s. The relationship of these various 16s components to each other is discussed.  相似文献   

20.
Purified replicative form (RF) and replicative intermediate (RI) prepared from Escherichia coli cells infected with the ribonucleic acid (RNA) bacteriophage R17 were denatured with dimethyl sulfoxide at 37 C or in aqueous solvents of low ionic strength at 97 C. Denaturation was demonstrated for RF and RI by an increase in specific infectivity and a striking change in the hyperchromicity curves after treatment. RI denaturation was also demonstrated by a shift in the buoyant density in Cs(2)SO(4) from 1.619 to the buoyant density of single-stranded R17 RNA (1.627). Analysis of the denatured RI hyperchromicity curves and the equilibrium distributions of denatured RI in Cs(2)SO(4) gradients revealed, however, a residual double-stranded component. Velocity sedimentation of denatured RI was performed, and the weight distribution of S values was calculated. From the known relation between molecular weight and S values, it was possible to transform the weight distribution into a number distribution of chain lengths. This distribution was compared with that predicted from the steady-state hypothesis for RI. Deviations from the predicted distribution may be due to the residual double-stranded component.  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