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.
Helix pomatia beta-haemocyanin was split into dissociation products by varying the pH and the ionic strength. The purity of the solution was checked in an ultracentrifuge. Two defined dissociation products were studied in solution by small-angle X-ray scattering. In Tris-HC1 buffer, pH 8.0 and ionic strength 1 M, the following parameters of the dissociation product (tenths) could be determined: molecular weight = 7 x 10(5), volume = 1350 nm3, radius of gyration = 9.0 nm, maximal distance = 28.3 nm, radius of the spherical subunits about 2.6 nm, number of the subunits approximately 19. Tris-HC1 buffer, pH 8.7 and ionic strength 0.01 M, yielded dissociation products (twentieths) with the following parameters: molecular weight = 3.5 x 10(5), volume = 635 nm3, radius of gyration = 7.5 nm, maximal distance = 21.9 nm, radius of the spherical subunits about 2.5 nm. With this information, the assumption that the larger fragment was double the smaller one and the latest biochemical and morphological results, theoretical scattering curves of models were calculated and compared with the experimental curves. Two models are suggested which argee well with the dissociation products in radius of gyration and scattering.  相似文献   

2.
Higher-order structures of chromatin in solution.   总被引:4,自引:0,他引:4  
Neutron scatter studies have been made on gently prepared chicken erythrocyte chromatin over a range of ionic strength. At low ionic strength the mass per unit length of the '10 nm nucleofilament corresponds to one nucleosome per 8--12 nm and a DNA packing ratio of between 6 and 9. From the contrast dependence of the cross-section radius of gyration of the nucleofilament the following parameters have been obtained; RgDNA' the cross-section radius of gyration (Rg) when DNA dominates the scatter; RgP, the cross-section Rg when protein dominates the scatter; Rc, the cross-section Rg at infinite contrast and alpha, the constant which describes the dependence of the cross-section Rg on contrast variation. From our understanding of the structure of the core particle, various arrangement of core particles in the nucleofilament have been tested. In models consistent with the above parameters the core particles are arranged edge-to-edge or with the faces of the core particles inclined to within 20 degrees to the axis of the nucleofilament. With increase of ionic strength the transition to the second-order chromatin structure has been followed. This gave the interesting result that above 20 microM NaCL or 0.4 mM MgCL2 the cross-section Rg increases abruptly to about 9 nm with a packing ratio of 0.2 nucleosome/mn and with further increase of ionic strength the Rg increases to 9.5 nm while the packing ratio increases threefold to 0.6 nucleosome/nm. This suggests a family of supercoils of nucleosomes which contract with increasing ionic strength. In its most contracted form the diameter of the hydrated supercoil has been found from the radial distribution function to be 34 nm. Models for the arrangements of core particles in the 34-nm supercoil are discussed.  相似文献   

3.
The formation of protein-polymer complexes was studied in an aqueous system using dynamic light scattering (DLS) and static light scattering (SLS) as the main experimental tools. Human serum albumin (HSA) was used as a protein and complexed with four representative water-soluble polymers: poly(N-isopropylacrylamide) (PNIPA), poly(ethylene glycol) (PEG), poly(vinyl pyrrolidone) (PVP), and poly(vinyl alcohol) (PVA). The first three molecular weights were within 420,000-540,000 and the last one was 270,000. The complexation was performed at 25 degrees C in 0.01 M NaCl solution adjusted to pH 3 with HCl as a function of mixing ratio (rm; molar ratio of polymer to HSA). From SLS experiments, we determined the molecular weight of the resulting complexes, from the value of which the number (nb) of bound proteins per polymer was estimated. It was found that each polymer forms an intrapolymer complex over a wide range of rm (1.2 > or = rm > or = 0.01). Then, a marked decrease in nb with increasing rm was found. Over the whole rm range, the HSA-PNIPA complex exhibited a large nb value, as compared with the other three complexes whose nb values at the same rm were close to one another. Both the hydrodynamic radius (Rh) by DLS and the radius of gyration (Rg) by SLS for the complexes of PNIPA, PVP, and PVA decreased and then reached a constant value as nb decreased with increasing rm. In the PEG system, however, there were a few changes in Rh and Rg with nb. The Rg/Rh ratio, as an indication of chain expansion, was found to increase with decreasing nb in the PNIPA system. The complexes of PVA and PVP displayed a similar tendency, although the magnitude of the increasing trend was smaller than that of the PNIPA complex. In contrast, the Rg/Rh ratio of the PEG complex hardly varied depending on nb. These results were discussed in connection with the differences of physicochemical properties among four water-soluble polymers.  相似文献   

4.
Small angle neutron scattering (SANS) results revealed that contrary to most reports C9 is not a globular protein. Its radius of gyration (Rg) at pH 8 and an ionic strength of 0.5 is 32.2 +/- 1.4 A increasing to 35 A at physiologic ionic strength. In contrast, C8, which has a 2.2-fold larger mass, has a similar Rg value [34.6 +/- 1.6 A]. Calibration plots of Rg vs. M(r) indicate that native C8 is a spherical protein whereas native C9 is elongated. From previous reports it was known that native C8 and C9 associate in solutions of low ionic strength. SANS results confirmed this observation but also demonstrated that C8-C9 heterodimers are already formed at physiologic ionic strength. The dimeric complex is globular [Rg = 40 +/- 0.8 A] indicating that the proteins associate side-by-side rather than end-to-end. In contrast, in presence of the drug Suramin, a potent inhibitor of the assembly of the C5b-9 complex, C9 forms a complex with twice the molecular mass that is still elongated (Rg = 48.8 +/- 0.8 A), suggesting that in this case the protein dimerizes end-to-end via a bridging Suramin molecule.  相似文献   

5.
Differential scanning calorimetry was performed to monitor the heat-induced changes that occur in the structural domain of lens alpha-crystallin. Circular dichroism and fluorescence also were used to resolve the controversial issue of the quaternary structure of alpha-crystallin. Based on the thermal behavior as monitored by these techniques, a model is proposed that can account for all previous data as well as the currently reported thermal data. The proposed model of native alpha-crystallin has a three-layer structure in which the inner layer (core) is a micelle containing 12 subunits arranged in cuboctahedral symmetry. The apolar region is directed inward constituting a hydrophobic core similar to a micelle and adding structural stability. A second layer of six subunits has a similar but not identical structure to the first layer, directing its apolar face toward the hydrophobic core. Thus, these two layers constitute a micelle-like structure with octahedral symmetry. The third layer adds more subunits for a total of not more than 24. Differential scanning calorimetry, circular dichroism, and fluorescence studies indicated that the inner two-layer structure of molecular mass 360 kDa is highly stable and is most likely of the alpha m form. The three-layer structure of the native protein, however, is rather unstable. At 35-45 degrees C the outer layer dissociates from the inner two layers, and at higher temperatures rapidly reassociates to a slightly modified two-layer structure with a stability similar to that of alpha m. The proposed model does not require any specific assembly of the alpha A and alpha B subunits in each layer, but the fluorescence results suggest that the native inner two layers probably contain mostly alpha A.  相似文献   

6.
The pigeon liver fatty acid synthetase complex (14 S) is dissociated in low ionic strength buffer containing dithiothreitol to form a half-molecular weight subunits (9 S) which are completely inactive for the synthesis of saturated fatty acids. The dithiothreitol-protected (reduced) subunits are rapidly reassociated and reactivated to form the active enzyme complex, not only by an increase in salt concentration but also by micromolar concentrations of NADP+ or NADPH. Increases in KCl or NADPH concentration result in an increase in the extent of reactivation (equilibrium) with no change in the over-all rate of the reaction or the half-life ofreactivation of the enzyme. The extent (equilibrium) of reactivation of the enzyme is the same in 0.2 M potassium phosphate buffer, pH 7.0; 0.2 M KCl in 5 mM Tris-35 mM glycine buffer, PH 8.3; and 50 muM NADP+ or NADPH in the Tris-glycine buffer. The extent and rate of reactivation of the enzyme is dependent not only on ionic strength and NADPH concentration, but also on pH and temperature. Reactivation with 0.2 M KCl is optimal between pH 7.3 and 8.5. At higher and lower pH values the rate and extent of reactivation are lowered. The rate and extent of reactivation are also decreased as the temperature is lowered below 10 degrees. At 0 degrees there is little reactivation of enzyme activity. However, in the presence of 0.2 M KCl containing 15 to 40% glycerol at 0 degrees, reactivation of the enzyme is about 50% complete. The rate of reactivation of enzyme in the presence of KCl or NADPH conforms to first order kinetics. This result suggests that the subunits first combine to form an inactive complex which is subsequently transformed to an enzymatically active complex. Evidence for the presence of inactive complex was obtained in experiments carried out in 0.2 M KCl at pH 6.0, and in 0.2 M KCl at pH 8.3, at both 6 and 3 degrees. Under these conditions the amount of complex observed upon ultracentrifugation was greater than expected from determinations of enzyme activity. The above findings suggest that ionic and hydrophobic interactions, and possibly the water structure surrounding the interacting sites, are of prime importance in reassociation and reactivation of enzyme. In addition, NADP+ and NADPH have very specific effects in bringing about reassociation and in maintaining the structural integrity of the multienzyme complex.  相似文献   

7.
Hsp16.5, isolated from the hyperthermophilic Archaea Methanococcus jannaschii, is a member of the small heat-shock protein family. Small Hsps have 12- to 42-kDa subunit sizes and have sequences that are conserved among all organisms. The recently determined crystal structure of Hsp16.5 indicates that it consists discretely of 24 identical subunits. Using fluorescence resonance energy transfer, we show that at temperatures above 60 degrees C, the subunits of MjHsp16.5 freely and reversibly exchange with a rate constant of exchange at 68 degrees C of 0.067 min(-1). The subunit exchange reactions were strongly temperature-dependent, similar to the exchange reactions of the alpha-crystallins. The exchange reaction was specific to MjHsp16.5 subunits, as other sHsps such as alpha-crystallin were not structurally compatible and could not integrate into the MjHsp16.5 oligomer. In addition, we demonstrate that at temperatures as high as 70 degrees C, MjHsp16.5 retains its multimeric structure and subunit organization. Using insulin and alpha-lactalbumin as model target proteins, we also show that MjHsp16.5 at 37 degrees C is a markedly inefficient chaperone compared with other sHsps with these substrates. The results of this study support the hypothesis that MjHsp16.5 has a dynamic quaternary structure at temperatures that are physiologically relevant to M. jannaschii.  相似文献   

8.
The rate of the 4 to 5 S estrogen-binding protein (EBP) in vitro transformation was measured by sucrose gradient centrifugation analysis. The temperature-activated 4 to 5 S EBP transformation is found to be highly reproducible without loss of [3H]estradiol-binding activity in a buffer containing an excess of [3H]estradiol, 40 mM Tris, 1 mM dithiothreitol, and 1 M urea at pH 7.4. The presence of [3H]estradiol is necessary for the 4 to 5 EBP transformation. A kinetic analysis of the 4 to 5 EBP transformation shows that it is a bimolecular reaction, the dimerization of the 4 S EBP with a second (similar or dissimilar) monomer or subunit. In buffers containing 0.4 M KCl the apparent second order rate constant is 2.3 plus or minus 0-2 times 10-7 M minus 1 min minus 1 at 28 degrees. The reaction is independent of the initial receptor concentration, suggesting that the 4 S EBP is dissociated into monomeric units in buffers of high ionic strength. In buffers without KCl or with 0.1 M KCl the apparent second order rate constant of receptor transformation increases with decreasing receptor concentration. This suggests that the 4 S EBP is associated weakly with another macromolecule (or macromolecules) in buffers of low ionic strength. The rate of 4 to 5 S EBP transformation shows a 200-fold increase between 0 and 35 degrees. The Arrhenius energy of activation is 21.3 kcal mol minus 1 in buffer without KCl and 19.1 kcal mol minus 1 in buffer with 0.4 M KCl. Following the temperature-activated dimerization, the avidity of binding between the 4 S EBP and its complementary subunit is increased, 0.4 M KCl can no longer cause dissociation, and the 5 S EBP dimer appears. This kinetic analysis indicates that the avidity of binding between the subunits of the estrogen receptor is modulated by estradiol, temperature, and ionic strength. We propose that these interactions of the estrogen receptor's subunits reflect conformational changes involved in receptor activation.  相似文献   

9.
The effects of pH on hyaluronate as observed by light scattering   总被引:1,自引:0,他引:1  
C E Reed  X Li  W F Reed 《Biopolymers》1989,28(11):1981-2000
Hyaluronate was investigated over a wide pH range, and at near zero and intermediate ionic strength, using dynamic and total intensity light scattering. Commercially obtained rooster comb hyaluronate was purified, and solutions were prepared in pure water by low-power bath ultrasonication and subsequent filtering. These solutions were of low polydispersity and appeared to contain single molecules of hyaluronate. Despite the absence of added electrolyte, these solutions yielded well-behaved Zimm plots. Increasing ionic strength and changing pH decreased radii of gyration and increased diffusion constants. Except for what appeared to be slow hydrolysis at either extreme of pH, molecular weights remained constant under all pH and ionic strength conditions. Under all solvent conditions investigated, diffusion coefficients increased with decreasing hyaluronate concentration. Unsonicated, lightly centrifuged solutions without added electrolyte were polydisperse, and their light scattering intensity was dominated by what appeared to be stable hyaluronate aggregates. The results are interpreted in terms of the polyelectrolyte properties of hyaluronate and its tendency to form stable entanglements, especially at low ionic strength. Previous light scattering studies in the literature on hyaluronate have shown widely varying results. The present article briefly reviews this literature and attempts to explain the variation among the previous results, emphasizing the Kuhn statistical segment length as an indicator of whether results are influenced by polydispersity or contaminants causing hyaluronate aggregation.  相似文献   

10.
Bioresponsive poly(amidoamine)s (PAA)s are currently under development as endosomolytic polymers for intracellular delivery of proteins and genes. Here for the first time, small-angle neutron scattering (SANS) is used to systematically investigate the pH-dependent conformational change of an endosomolytic polymer, the PAA ISA 23. The radius of gyration of the ISA23 was determined as a function of pH and counterion, the aim being to correlate changes in polymer conformation with membrane activity assessed using a rat red blood cell haemolysis assay. With decreasing pH, the ISA23 radius of gyration increased to a maximum (R(g) approximately 80 A) around pH = 3, before subsequently decreasing once more. At high pH and therefore high ionic strengths, the polymer is negatively charged and adopts a rather compact structure (R(g) approximately 20 A), presumably with the dissociated carboxylic groups on the exterior of the polymer coil. At low pH, the coil again collapses (R(g) < 20 A), presumably due to the effects of the high ionic strength. It is concluded that the nature of the salt form has no direct bearing on the size of the polymer coil, but it does indirectly determine the prevailing pH and, hence, polymer conformation. Pulsed-gradient spin-echo NMR measurements were in good agreement with the SANS estimates of the radius of gyration, although ISA23 polydispersity does complicate the data interpretation/comparison. These results support the proposed mode of action of PAAs, namely a coil expansion on passing from a neutral pH (extracellular) to an acidic pH (endosomal and lysosomal) environments. The results do, however, suggest that the charge on the polymer shows a closer correlation with the haemolysis activity rather than the polymer conformation.  相似文献   

11.
The binding of labeled hyaluronate to the surface of Simian virus 40-transformed 3T3 cells was studied as a function of 1) pH, 2) ionic strength, 3) temperature, and 4) molecular weight of the hyaluronate. Binding occurred over a wide range of pH values with optima at pH 7 and at less than pH 4. Binding at low pH was eliminated at high ionic strength whereas that at physiological pH was enhanced, with a maximum at 0.5 M NaCl. The enhancement of binding at pH 7 was reversible and independent of the particular salt used. Scatchard plot analysis showed that increasing the ionic strength resulted in both a decrease in the dissociation constant (Kd) and an increase in the amount bound at saturation (Bmax). Temperature also influenced the binding of hyaluronate to the cell surface. The amount bound at low temperatures (0 degrees C) was 3 to 5 times that bound at high temperatures (40 degrees C) with a sharp transition occurring at 18 degrees C, the temperature of phase transition of the plasma membrane. The temperature effect was primarily a change in the Bmax and was reversible. Finally the molecular weight of the ligand influenced the binding. High molecular weight preparations of hyaluronate had a higher binding affinity (lower Kd) and a lower Bmax than did smaller molecular weight preparations.  相似文献   

12.
Changes in the conformational state of chromatin core particles from chicken erythrocytes were studied by both immunochemical and biophysical methods as a function of pH and ionic strength. When the pH of core particles in a solution of ionic strength 3, 60 or 220 mM was lowered from pH 7.5, a sharp transition in the circular dichroism spectrum of DNA monitored between 320 and 260 nm was observed at pH 6.65. This change in DNA ellipticity was totally reversible. Binding to core particles of antibodies specific for histones H2B, H2A, H3 and for the IRGERA (synthetic C-terminal) peptide of H3 was used to follow changes in histone antigenicity. Binding was studied in the pH range 7.5-5.35, and at ionic strength of 60 and 220 mM. A change in reactivity of some histone epitopes was observed around pH 6.2–6.5. However, the changes observed by circular dichroism and antibody binding pertain to different components of chromatin subunits and they probably reflect independent phenomena. The alteration in accessibility of these determinants at the surface of core particles was completely reversible and was dependent on ionic strength. The conformation changes in core particles occurring near physiological ionic strength and pH may reflect dynamic changes in chromatin structure that possess functional significance.  相似文献   

13.
The ionization of 4-nitroimidazole to 4-nitroimidazolate was investigated as a function of ionic strength. The apparent pKa varies from 8.99 to 9.50 between 0.001 and 1.0 M ionic strength, respectively, at 25 degrees C. The ionic strength dependence of this ionization is anomalous. The binding of 4-nitroimidazole by horse metmyoglobin was studied between pH 5.0 and 11.5 and as a function of ionic strength between 0.01 and 1.0 M. The association rate constant is pH-dependent, varying from 24 M(-1)s(-1) at pH 5 to a maximum value of 280 M(-1)s(-1) at pH 9.5 and then decreasing to 10 M(-1)s(-1) at pH 11.5 in 0.1 M ionic strength buffers. The dissociation rate constant has a much smaller pH dependence, varying from 0.082 s(-1) at low pH to 0.035 s(-1) at high pH, with an apparent pKa of 6.5. The binding affinity of 4-nitroimidazole to horse metmyoglobin is about 2.5 orders of magnitude stronger than that for imidazole and this increased affinity is attributed to the much slower dissociation rate for 4-nitroimidazole compared to that of imidazole. Although the ionic strength dependence of the binding rate is small and secondary kinetic salt effects can account for the ionic strength dependence of the association rate constant, the pH dependence of the rate constants and microscopic reversibility arguments indicate that the anionic form of the ligand binds more rapidly to all forms of metmyoglobin than does the neutral form of the ligand. However, the spectrum of the complex is similar to model complexes involving neutral imidazole and not imidazolate. The latter observation suggests that the initial metmyoglobin/4-nitroimidazolate complex rapidly binds a proton and the neutral form of the bound ligand is stabilized, probably through hydrogen binding with the distal histidine.  相似文献   

14.
Troponin T kinase utilizes ATP and GTP as substrates for the protein kinase reaction. When phosvitin is used as substrate, the enzyme activity increases with an increase in the ionic strength up to 0.2-0.3 M KCl. The pH optimum for the enzyme lies at 8-9. Ultracentrifugation in sucrose density gradient showed that the sedimentation coefficient of troponin T kinase is 9.5 S. The Stocks radius determined by gel-filtration is equal to 49 A. The molecular weight of the enzyme calculated from the given values of the Stocks radius and sedimentation coefficient is 184,000. This is indicative of the oligomeric structure of the enzyme and suggests that the stoicheiometry of the enzyme subunits having mol. weights of 50,000, 46,000 and 31,000 is other than 1:1:1. Troponin T kinase is capable of autophosphorylation; the phosphorylation process involves only the protein with mol. weight of 31,000. The data obtained suggest that troponin T kinase can be referred to casein kinases of G type and may participate in translation processes.  相似文献   

15.
In order to clarify changes in the structure and surface properties of photoactive yellow protein (PYP) upon light absorption, the spectroscopic properties and solution structure of its photo-intermediate (PYP(M)) were examined in the presence of various anions. At identical ionic strengths, citrate slowed the decay rate of PYP(M) more than acetate. Although the absorption spectrum in the dark was not affected by organic anions, citrate induced a 5-nm blue shift of the absorption maximum for PYP(M). Solution X-ray scattering experiments indicated that the radius of gyration (Rg) and apparent molecular weight in the dark were constant in all buffer systems. However, the Rg of PYP(M) in citrate buffer at high concentration was 16.2 (+/-0.2) A, while the Rg of PYP(M) in acetate buffer was 15.6 (+/-0.2) A. The apparent molecular weight increased 7% upon PYP(M) formation in citrate buffer at high concentration compared to other conditions. These results suggest that citrate molecules specifically bind to PYP(M). A cluster of basic amino acid residues with a hydrogen bond donor would be exposed upon PYP(M) formation and responsible for the specific binding of citrate.  相似文献   

16.
The association behavior of beta-lactoglobulin has been studied by small-angle neutron scattering as a function of protein concentration, temperature, pH, and NaCl concentration of the solution. By indirect Fourier transformation of the spectra, pair-distance distribution functions for the various samples were obtained. These functions provided information on the maximum size, the weight-averaged molecular mass, and the z-averaged radius of gyration of the beta-lactoglobulin particles. At room temperature and pH values below 4 and above 5.2 the protein consists predominantly of monomers and dimers, consistent with literature. In these pH regimes the formation of dimers is favored upon increasing ionic strength and decreasing protein charge (pH values closer to the isoelectric point of the protein). Around pH 4.7, larger oligomeric structures are formed, enhanced by a decrease in temperature and a decrease in ionic strength. beta-Lactoglobulin A associates more strongly than beta-lactoglobulin B. Surprisingly, at pH 6.9 larger structures than dimers seem to be formed at high protein concentrations (> 30 mg mL-1).  相似文献   

17.
The self-association of human spectrin has been studied by means of sedimentation equilibrium in the analytical ultracentrifuge at pH 7.5 and over a range of ionic strength from 0.009 to 1.0 M. Increasing ionic strength above 0.1 M reduces the equilibrium constants for all of the measurable steps in the self-association reaction. These results support the concept of charge-charge interactions stabilizing the tetramer and higher oligomers with respect to the heterodimer. In addition, increasing ionic strength brought about a dissociation of the heterodimer to component polypeptide chains. Dissociation to the heterodimers is also enhanced with a decrease in ionic strength below 0.05 M. This low ionic strength-dependent dissociation is consistent with generalised electrostatic repulsion; however, this effect also correlates with some loss of alpha-helical content as revealed by circular dichroism. The secondary, tertiary and quaternary structures may all be partially disrupted by electrostatic free energy at low ionic strength.  相似文献   

18.
Lumbricus terrestris HbO2 and HbCO dissociated below pH 5.0; a time-dependent alteration to the met form occurred at pH less than 5 and pH less than 4.5, respectively. The extent of dissociation was unaffected by alkaline earth cations but was decreased by an increase in ionic strength. HbO2 and HbCO exposed to pH 4.0-4.8 were centrifuged to obtain the undissociated pellet (P1) and dissociated supernatant (S1) fractions. S1 was reassociated at pH 7.0 by dialysis against various buffers and then centrifuged to obtain the reassociated pellet (P2) and unreassociated supernatant (S2) fractions. Reassociation was possible only if S1 was dialyzed against water prior to return to neutral pH; otherwise precipitation occurred starting at about pH 5.3. The extent of reassociation varied from about 40 to 80%, was usually higher for HbCO than HbO2, and was unaffected by an increase in ionic strength or by Ca(II). Gel filtration of P2 on Sephacryl S-300 at neutral pH gave one peak IaR, eluting at a slightly greater volume than the native Hb; S1 and S2 gave in addition, three peaks, Ib (200 kDa), II (65 kDa), and III (18 kDa). Sodium dodecyl sulfate-polyacrylamide gel electrophoresis showed that P2 was slightly deficient in subunit M relative to the Hb, that Ib was deficient in subunits D1 and D2 and that II and III consisted of subunits D1 + D2 + T and subunit M, respectively. Scanning transmission electron microscopy of P2 showed that it was smaller than the native hemoglobin: 25 nm in diameter and 16 nm in height, instead of 30 X 20 nm. Comparison of the results of the dissociations of Lumbricus Hb at alkaline pH (Kapp, O. H., Polidori, G., Mainwaring, M., Crewe, A. V., Vinogradov, S. N. (1984) J. Biol. Chem. 259, 628-639) with those obtained in this study suggested that the Hb quaternary structure was not multimeric and that an alternative model had to be considered. In the proposed model it is assumed that subunits D1 and D2 form a scaffolding or "bracelet," decorated with 12 complexes of M and T subunits.  相似文献   

19.
The radii of gyration of unfolded apo-cytochrome C at pH 2.3 have been determined in three conditions: (i) 20 mM sodium phosphate buffer; (ii) 0.25 M NaCl; and (iii) 6.65 M GuHCl by small-angle X-ray scattering, and (iii) from translational diffusion coefficients measured by dynamic light scattering. The radius of gyration of the unfolded protein chain depends remarkably on the quality of the solvent, decreasing in the order 20 mM sodium phosphate greater than 6.65 M GuHCl greater than 0.25 M NaCl. The value of the radius of gyration in 0.25 M NaCl and also the value estimated for infinite ionic strength are close to the value predicted theoretically for the theta-point. This means that water in the absence of electrostatic interactions is a poor solvent for an unfolded protein while 6.65 M GuHCl is a better solvent.  相似文献   

20.
J Loscalzo  M Fisch  R I Handin 《Biochemistry》1985,24(16):4468-4475
The reversible association of protomers of von Willebrand protein (vWF) was studied in order to analyze the forces and mechanism of vWF polymer assembly. At concentrations of vWF found in plasma (approximately 16 micrograms/mL), disulfide bond reduction with 50 mM 2-mercaptoethanol (2-ME) markedly reduced both vWF activity, as measured by ristocetin-dependent platelet agglutination, and average polymer size (Rh, the mean hydrodynamic radius) in solution, as determined by quasi-elastic light scattering (QLS) and by gel filtration chromatography. With increasing vWF concentration, activity and Rh increased despite reduction of interprotomer disulfide bonds. Changes in temperature after 2-ME treatment produced reversible changes in activity and Rh. Varying the total vWF concentration at any given temperature after 2-ME treatment changed Rh in a consistent and predictable fashion, so that estimates of the dissociation constant for vWF protomer-polymer equilibrium were obtained: Kd5 degrees C = 0.77 micrograms/mL, Kd25 degrees C = 2.4 micrograms/mL, and Kd37 degrees C = 7.7 micrograms/mL, where under the conditions of reduction presented here, the basic protomer of vWF is a dimer. Increasing ionic strength after 2-ME treatment with 1 M KCl did not change Rh, while approximately 100 microM sodium dodecyl sulfate (SDS) or approximately 300 microM sodium deoxycholate (DOC) reduced both Rh and activity compared with those of unreduced polymer. These data show that disulfide bonds are necessary to maintain vWF polymer size and activity at plasma concentrations but that noncovalent forces of association can maintain vWF polymer size and activity at higher concentrations. These forces of association may be important for polymer assembly during intracellular synthesis of vWF.  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