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.
Sephadex G-10 gel sieving chromatography, Jones-Dole viscosity B coefficients, and solution neutron and X-ray diffraction are used to show that small ions of high charge density (e.g., sulfate, phosphate, the carboxylate, sodium, and fluoride) are strongly hydrated (kosmotropes) whereas large monovalent ions of low charge density (e.g., ammonium, chloride, potassium, and the positively charged amino acid side chains) are weakly hydrated (chaotropes). The heats of solution of the crystalline alkali halides are then used to show that only oppositely charged ions of equal water affinity spontaneously form inner sphere ion pairs, and that this controls ion binding to proteins. The net charge on a protein is a major determinant of its solubility. Finally, the surface potential difference and surface tension at an air-salt solution interface are used to generate a simple model for how ions affect protein stability and solubility through indirect interactions at the protein-solution interface. A few comments about small neutral osmolytes are also included.  相似文献   

2.
Large sign-alternating charge clusters formed by the charged side groups of amino acid residues and N- and C-terminal groups were found in the majority of considered globular proteins, namely 235 in a total of 274 protein structures, i.e. 85.8%. The clusters were determined by the criteria proposed earlier: charged groups were included in the cluster if their charged N and O atoms were located at distances between 2.4 and 7.0 A. The set of selected proteins consisted of known non-homologous protein structures from the Protein Data Bank with a resolution less than or equal to 2.5 A and pair sequence similarity less than 25%. Molecular masses of the proteins were from 5.5 to 91.5 kDa and protein chain length from 50 to 830 residues. The distribution of charged groups on the protein surface between isolated charged groups, small clusters with two and three groups, and large clusters with four or more groups were found to be approximately similar making 33, 35 and 32% of the total amount of protein charged groups, respectively. The large sign-alternating charge clusters with four or more charged groups were studied in greater detail. The amount of such clusters depends on the protein chain length. The small proteins contain 1-3 clusters while the large proteins display 4-6 or more clusters. On average, 1.5 clusters per each 100 residues were observed. In contrast with this, the size of a cluster, i.e. the number of charged groups inside a cluster, does not depend on the protein molecular mass, and large clusters are observed for proteins from a range of molecular masses. Clusters consisting of four to six charged groups occur most frequently, although extra large clusters are also often revealed. We can conclude that sign-alternating charge clusters are a common feature of the protein surface of globular protein. They are suggested to play a general functional role as a local polar factor of protein surface.  相似文献   

3.
A series of charge-modified thaumatins with different values of surface charge were partitioned in aqueous two-phase systems (ATPS) to study the effect of surface charge as a single property on partitioning. Electrophoretic mobility of the proteins in titration curves was used as a measure of surface charge. Four modified proteins derived from thaumatin with the following values of isoelectric point: 8.70, 8.15, 5.60, and 4.50 were used for partitioning. The resolution of the systems in terms of protein surface charge was calculated. Partitioning of modified thaumatins in PEG 4000/dextran systems with phosphate buffer, Tris buffer, NaCl, KCl, and sulfate salts was carried out. Among the sulfate salts tested, the addition of 50 mM Li(2)SO(4) to the system buffered with phosphate gave the highest value of resolution for differences in surface protein charge (RSPC). It shows a decrease in the value of K (partition coefficient) with an increase in the protein's charge. The addition of 100 mM KCl to the system promoted the opposite effect on the RSPC value. Charge-modified proteins were partitioned in PEG/salt systems to investigate the ability of these systems for resolving differences in surface charge. The PEG/citrate system seemed to have almost no ability for resolving proteins on the basis of surface charge differences; PEG/phosphate systems had some capability for resolving differently charged proteins. The more negative proteins tended to have higher values of K than the more positively charged fractions. The use of charge-modified proteins allowed the investigation of the effect of protein surface charge on partitioning in aqueous two-phase systems independently from other protein parameters as they were prepared from a common parent protein thaumatin. This technique provides an interesting novel tool to investigate the effect of protein surface charge on partitioning in ATPS taking protein charge as an independent parameter. (c) 1996 John Wiley & Sons, Inc.  相似文献   

4.
BACKGROUND: The traditional picture of charged amino acids in globular proteins is that they are almost exclusively on the outside exposed to the solvent. Buried charges, when they do occur, are assumed to play an essential role in catalysis and ligand binding, or in stabilizing structure as, for instance, helix caps. RESULTS: By analyzing the amount and distribution of buried charged surface and charges in proteins over a broad range of protein sizes, we show that buried charge is much more common than is generally believed. We also show that the amount of buried charge rises with protein size in a manner which differs from other types of surfaces, especially aromatic and polar uncharged surfaces. In large proteins such as hemocyanin, 35% of all charges are greater than 75% buried. Furthermore, at all sizes few charged groups are fully exposed. As an experimental test, we show that replacement of the buried D178 of muconate lactonizing enzyme by N stabilizes the enzyme by 4.2 degrees C without any change in crystallographic structure. In addition, free energy calculations of stability support the experimental results. CONCLUSIONS: Nature may use charge burial to reduce protein stability; not all buried charges are fully stabilized by a prearranged protein environment. Consistent with this view, thermophilic proteins often have less buried charge. Modifying the amount of buried charge at carefully chosen sites may thus provide a general route for changing the thermophilicity or psychrophilicity of proteins.  相似文献   

5.
The enhancement of protein thermostability is an important issue for both basic science and biotechnology purposes. We have developed a thermostability criterion for a protein in terms of a quasi-electric dipole moment (contributed by its charged residues) defined for an electric charge distribution whose total charge is not zero. It was found that minimization of the modulus of this dipole moment increased its thermal stability, as demonstrated by surveying these values in pairs of mesostable-thermostable homologous proteins and in mutations described in the literature. The analysis of these observations provides criteria for thermostabilization of a protein, by computing its dipole profile. This profile is obtained by direct substitution of each amino acid of the sequence by either a positive, negative or neutral amino acid, followed by a recalculation of the dipole moment. As an experimental example, these criteria were applied to a beta-glucanase to enhance its thermal stability.  相似文献   

6.
7.
A purified protein fraction from the proteolipids of human brain myelin was recombined with different lipids either in aqueous buffer or in a chloroform-methanol-water (10:5:1, v/v/v) mixture. It was found that under both conditions it binds strongly to phospholipids irrespective of surface charge, the presence of cholesterol or double bonds on the fatty acyl chains. The buoyant density of the resulting lipoprotein membranes is intermediate to that of pure lipids, and proteins. The lipoproteins formed by either of these methods were observed by either freeze-fracture or negative stain electron-microscopy. The overall morphology was similar to that of pure phospholipids, showing large closed multilamellar vesicles. The presence of the protein was detected by the appearance of intramembrane particles in freeze-fracture. The addition of the N-2 protein generally increases the permeability vesicles to 22-Na-+ by 2-3 orders of magnitude depending on the concentration. The presence of calcium in the aqueous medium further increases the Na-+ efflux through negatively charged vesicles. Changes in lipid composition, surface charge, cholesterol, etc., have no appreciable influence on the effect of the protein. Differential scanning calorimetry indicates that the presence of small amounts of N-2 have no effect on the lipid phase transition from solid to liquid crystalline. As the amount of protein bound to the phospholipid increases, the enthalpy of the transition decreases, the main endothermic peak broadens, but there is no change on the midpoint temperature. Membranes containing 50% by weight of protein still show a transition with an enthalpy approximately one half that of the original lipid.  相似文献   

8.
Three electromorphs of iron superoxide dismutase (FeSOD) were identified among 29 Helicobacter pylori isolates by native gel electrophoresis and activity staining. The electromorphs designated isoforms A, B, and C are characterized by slow, intermediate and fast electrophoretic migration, respectively, which was not observed under denaturing conditions. The isoforms were not associated with virulence determinants and with the outcome of disease. Sequence analysis of the sodB gene in strains producing different FeSOD isoforms and comparison of deduced protein sequences revealed that differences in the electric migration behavior are associated with exchange of charged amino acids, suggesting that faster migration is caused by a more negative total charge of the proteins. Electrophoretic migration of native FeSOD was not influenced by changes in the iron cofactor concentration, oxidative stress, and different media, indicating that FeSOD isoforms represent stable strain-specific markers.  相似文献   

9.
The sign and magnitude of the surface charge of liposomes containingelectrostatically neutral lecithin and cholesterol was alteredby incremental additions of dicetyl phosphate or stearylamine.Such liposomes instantaneously bound authentic proteins at 0°Conly when they had electrostatically opposite charges; 1 M NaClinhibited the binding. The amount of protein bound was dependentupon the concentration of protein and the charge of liposomes.Phytochrome in a crude extract of etiolated pea (Pisum sativumcv. Alaska) shoots could bind equally well to liposomes witheither positive or negative charges irrespective of PR and PFRboth of which showed no spectral distortion. Both PR and PFRof purified pea phytochrome bound entirely to positively chargedliposomes but partially to negatively charged ones. In thisassociation both PR and PFR became pelletable at similar rates.Absorption spectra of liposome-bound PR showed a small blueshift and then a crucial spectral distortion after red-lightirradiation. (Received October 22, 1980; Accepted January 22, 1981)  相似文献   

10.
The interaction of nuclear proteins of protamine and histone with neutral and charged BLM was studied. Anion and cation detergents were used to create the surface charge. The surface density of charges in BLM was comparable with that in biomembranes. Protamine and histone increased the electroconductivity of negatively charged BLM for anions and cations correspondingly. It is suggested that the surface charge of the membrane may influence the ion transport directly and indirectly due to the interaction of the membrane structures with charged proteins present in the surrounding medium.  相似文献   

11.
The molecular details of how lipids influence final topological organization of membrane proteins are not well understood. Here, we present evidence that final topology is influenced by lipid–protein interactions most likely outside of the translocon. The N-terminal half of Escherichia coli lactose permease (LacY) is inverted with respect to the C-terminal half and the membrane bilayer when assembled in mutants lacking phosphatidylethanolamine and containing only negatively charged phospholipids. We demonstrate that inversion is dependent on interactions between the net charge of the cytoplasmic surface of the N-terminal bundle and the negative charge density of the membrane bilayer surface. A transmembrane domain, acting as a molecular hinge between the two halves of the protein, must also exit from the membrane for inversion to occur. Phosphatidylethanolamine dampens the translocation potential of negative residues in favor of the cytoplasmic retention potential of positive residues, thus explaining the dominance of positive over negative amino acids as co- or post-translational topological determinants.  相似文献   

12.
13.
Invertase was ionically bound to the poly(ethylene-vinyl alcohol) membrane surface modified with two aminoacetals with different molecular length, 2-dimethyl-aminoacetoaldehyde dimethylacetal (AAA) and 3-(N, N-dimethylamino-n-propanediamine) propionaldehyde dimethylacetal (APA). Immobilization conditions were determined with respect to enzyme concentration in solution, pH value, ionic strength in immobilization solution, and immobilization time. Various properties of immobilized invertase were evaluated, and thermal stability was found especially to be improved by immobilization. The apparent Michaelis constant, K(m), was smaller for invertase bound by APA with longer molecular lengths than for invertase bound by AAA. We attempted to bind glucoamylase of Rhizopus delemar origin in the same way. The amount and activity of immobilized glucoamylase were much less than of invertase.  相似文献   

14.
Correlations to describe the effect of surface hydrophobicity and charge of proteins with their partition coefficient in aqueous two-phase systems were investigated. Polyethylene glycol (PEG) 4000/phosphate, sulfate, citrate, and dextran systems in the presence of low (0.6% w/w) and high (8.8% w/w) levels of NaCl were selected for a systematic study of 12 proteins. The surface hydrophobicity of the proteins was measured by ammonium sulfate precipitation as the inverse of their solubility. The hydrophobicity values measured correlated well with the partition coefficients, K, obtained in the PEG/salt systems at high concentration of NaCl (r = 0.92-0.93). In PEG/citrate systems the partition coefficient correlated well with protein hydrophobicity at low and high concentrations of NaCl (r = 0.81 and 0.93, respectively). The PEG/citrate system also had a higher hydrophobic resolution than other systems to exploit differences in the protein's hydrophobicity. The surface charge and charge density of the proteins was determined over a range of pH (3-9) by electrophoretic titration curves; PEG/salt systems did not discriminate well between proteins of different charge or charge density. In the absence of NaCl, K decreased slightly with increased positive charge. At high NaCl concentration, K increased as a function of positive charge. This suggested that the PEG-rich top phase became more negative as the concentration of NaCl in the systems increased and, therefore, attracted the positively charged proteins. The effect of charge was more important in PEG/dextran systems at low concentrations of NaCl. In the PEG/dextran systems at lower concentration of NaCl, molecular weight appeared to be the prime determinant of partition, whereas no clear effect of molecular weight could be found in PEG/salt systems.  相似文献   

15.
The adsorption free energy of charged proteins on mixed membranes, containing varying amounts of (oppositely) charged lipids, is calculated based on a mean-field free energy expression that accounts explicitly for the ability of the lipids to demix locally, and for lateral interactions between the adsorbed proteins. Minimization of this free energy functional yields the familiar nonlinear Poisson-Boltzmann equation and the boundary condition at the membrane surface that allows for lipid charge rearrangement. These two self-consistent equations are solved simultaneously. The proteins are modeled as uniformly charged spheres and the (bare) membrane as an ideal two-dimensional binary mixture of charged and neutral lipids. Substantial variations in the lipid charge density profiles are found when highly charged proteins adsorb on weakly charged membranes; the lipids, at a certain demixing entropy penalty, adjust their concentration in the vicinity of the adsorbed protein to achieve optimal charge matching. Lateral repulsive interactions between the adsorbed proteins affect the lipid modulation profile and, at high densities, result in substantial lowering of the binding energy. Adsorption isotherms demonstrating the importance of lipid mobility and protein-protein interactions are calculated using an adsorption equation with a coverage-dependent binding constant. Typically, at bulk-surface equilibrium (i.e., when the membrane surface is "saturated" by adsorbed proteins), the membrane charges are "overcompensated" by the protein charges, because only about half of the protein charges (those on the hemispheres facing the membrane) are involved in charge neutralization. Finally, it is argued that the formation of lipid-protein domains may be enhanced by electrostatic adsorption of proteins, but its origin (e.g., elastic deformations associated with lipid demixing) is not purely electrostatic.  相似文献   

16.
Porins are integral proteins of the outer membranes of gram-negative bacteria. In membranes, they exist as homotrimers and the L2 loops contribute to their stability. Comparison of OmpC porins of the Yersinia pseudotuberculosis complex with other enterobacterial porins demonstrated L2 loop length diversity, which is caused by varying numbers of dipeptide/tripeptide repeats. The OmpC porins are highly homologous to each other, and they can be subdivided into five isoforms based on their L2 loop structure. Optical spectroscopy and SDS-PAGE experiments revealed that particularities of the L2 loops affected the structure and thermal stability of the porins. Thermal denaturation studies showed that porins with shorter loops, compared to porins with longer loops, had more stable tertiary and less stable secondary and quaternary structures. According to our comparative modeling results, the L2 loops differ in their structure by adopting different spatial positions and forming different polar bonds with a neighbor monomer. The replacement of asparagine with arginine at the C-terminus of the L2 loop shifts the loop upwards and causes the loss of contacts with the arginine clusters within the pores. The increase in the length of these loops ensures that they shift down toward the pore and restore their contacts with arginines on the channel wall, as is the case in classical nonspecific porins. Despite the fact that the surface charge density varies considerably among the OmpC porins, the L2 loops form a typical negatively charged region in the center of the trimer.  相似文献   

17.
Investigation on the volume properties of protein hydration layers is reported. Presented results are based on combination of Monte Carlo modeling and available experimental data. Six globular proteins with known data are chosen for analysis. Analyzing the model and the experimental results we found that water molecules bound to proteins by hydrogen bond are preferentially located at the places with local depressions on the protein surface. Consequently, the hydration level is not strictly proportional to the area of charged and polar surfaces, but also depends on the shape of the molecular surface. The thickness of the thermal volume layer as calculated in the framework of the scaled particle theory is 0.6-0.65 A for chosen proteins. The obtained value is significantly lower than that presented for proteins in earlier papers (where proportionality between the hydration level and the area of charged and polar surfaces was assumed), but is close to the value published for small solute molecules. Discussion including the influence of protein size and the thermal motion of the surface is presented.  相似文献   

18.
Anion and cation effects on the structural stability of lysozyme were investigated using differential scanning calorimetry. At low concentrations (<5 mM) anions and cations alter the stability of lysozyme but they do not follow the Hofmeister (or inverse Hofmeister) series. At higher concentrations protein stabilization follows the well‐established Hofmeister series. Our hypothesis is that there are three mechanisms at work. At low concentrations the anions interact with charged side chains where the presence of the ion can alter the structural stability of the protein. At higher concentrations the low charge density anions perchlorate and iodide interact weakly with the protein. Their presence however reduces the Gibbs free energy required to hydrate the core of the protein that is exposed during unfolding therefore destabilizing the structure. At higher concentrations the high charge density anions phosphate and sulfate compete for water with the protein as it unfolds increasing the Gibbs free energy required to hydrate the newly exposed core of the protein therefore stabilizing the structure.  相似文献   

19.
The binding of sodium dodecyl sulfate to coat protein subunits of cucumber green mottle mosaic virus and tobacco mosaic virus was studied by equilibrium dialysis. The amount of dodecyl sulfate bound to the cucumber virus protein in 0.1 m phosphate buffer (pH 7.2) was found to be 1.55 g/g, which was the same value as that obtained with the tobacco virus protein. The presence of 8 m urea markedly decreased the degree of binding of dodecyl sulfate to the proteins. The amount of binding to the cucumber virus protein was reduced to 0.56 g/g, and that to the tobacco virus protein decreased to 0.8 g/g. The net charges of both proteins were negative at neutral pH and the amount of negative charge of the cucumber virus protein, obtained from the potentiometric titration curves, was larger than that of the tobacco virus protein, either in the native state or in the denatured state. In dodecyl sulfate/polyacrylamide gel electrophoresis the cucumber virus protein migrated faster than the tobacco virus protein. On the other hand, in the presence of 8 m urea, the electrophoretic migration rate of the cucumber virus protein was equal to that of the tobacco virus protein. Sedimentation equilibrium experiments in 6 m guanidinium chloride gave molecular weights of 17,700 and 17,200 for the tobacco mosaic virus and the cucumber virus proteins, respectively. These results suggest that the effective negative charge density of the cucumber virus protein-dodecyl sulfate complex is higher than that of the tobacco virus proteindodecyl sulfate complex in 0.1% dodecyl sulfate solution. The conformation of both proteins was investigated by circular dichroism measurements. Both proteins have a slightly higher degree of α-helix content in dodecyl sulfate solution than in the native state. The addition of 8 m urea to both proteins while in this solution induced a change in conformation to one having a much smaller degree of ordered structure, although the change in the cucumber virus protein was more intense than that in the tobacco virus protein.  相似文献   

20.
Three soluble invertase (EC 3.2.1.26) isoforms from Easter lily ( Lilium longiflorum Thunb. cv. Nellie White) flower buds were purified to apparent homogeneity. Non‐denaturing PAGE showed one band for all three invertases that corresponded to the invertase activity. SDS‐PAGE of purified invertase I gave a single band at 78 kDa, whereas invertases II and III gave three bands at 54, 52 and 24 kDa. Antibodies against tomato fruit acid invertase and Urtica dioica leaf acid invertase recognized all three invertase isoforms, whereas antibodies against wheat coleoptile acid invertase recognized only 56‐ and 54‐kDa bands of invertases II and III. Antibodies against wheat coleoptile invertase recognized the 54‐ and 52‐kDa proteins from crude extracts of all flower organs, and a 72‐kDa protein in both leaf and bulb scale extracts. All three invertases bound to Con‐A peroxidase. Deglycosylation of invertase I with glycopeptidase F was complete and resulted in a peptide of 75 kDa. Invertases II and III were deglycosylated partially by glycopeptidase F and resulted in proteins of 53, 51, 50 and 22 kDa. Invertase I was localized only in anther and filament, whereas the other two isoforms were present in all flower organs.  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