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 impact of an extensive, uniform and hydrophobic protein surface on the behavior of the surrounding solvent is investigated. In particular, focus is placed on the possible enhancement of the structure of water at the interface, one model for the hydrophobic effect. Solvent residence times and radial distribution functions are analyzed around three types of atomic sites (methyl, polar, and positively charged sites) in 1 ns molecular dynamics simulations of the α-helical polypeptide SP-C in water, in methanol and in chloroform. For comparison, water residence times at positively and negatively charged sites are obtained from a simulation of a highly charged α-helical polypeptide from the protein titin in water. In the simulations the structure of water is not enhanced at the hydrophobic protein surface, but instead is disrupted and devoid of positional correlation beyond the first solvation sphere. Comparing solvents of different polarity, no clear trend toward the most polar solvent being more ordered is found. In addition, comparison of the water residence times at nonpolar, polar, positively charged, or negatively charged sites on the surface of SP-C or titin does not reveal pronounced or definite differences. It is shown, however, that the local environment may considerably affect solvent residence times. The implications of this work for the interpretation of the hydrophobic effect are discussed. Proteins 27:395–404, 1997. © 1997 Wiley-Liss, Inc.  相似文献   

2.
我们计算了氨基酸的溶剂可接近面积,局域偶极矩及表面电场.结果显示了亲、疏水基团与水相互作用具有不同的微观本质,由此提出了新的估计氨基酸溶剂化能的方法.  相似文献   

3.
The calculation of protein densities from atomic coordinates is not straight-forward and requires very careful attention to the determination of the protein-solvent boundary. Interior densities are more readily obtained and are in reasonable agreement with those estimated from solvent accessibility studies. The interior of globular proteins has very significant density inhomogenities on a scale of 100--1000 A3. The interior densities range from less than 0.5 g/cm3 to over 3 g/cm3. The low local densities are primarily associated with clusters of nonpolar sidechains while the high local density regions arise from the protein backbone secondary structures: helices and beta sheets. We show a rough correlation between local density and local polarity.  相似文献   

4.
The collapse transition of lattice protein-like heteropolymers has been studied by means of the Monte Carlo method. The protein model has been reduced to the α-carbon trace restricted to a high coordination lattice. The sequences of model heteropolymers contain two types of mers: hydrophobic/nonpolar (H) and hydrophilic/polar (P). Interactions of HH and PP pairs were assumed to be negative (weaker attractions of PP pairs) while the contact energy for HP pairs was equal to zero. All sequence-specific short-range interactions have been neglected in the present studies. It has been found that homopolymeric chains undergo a smooth collapse transition to a dense globular state. The globule lacks any signatures of local ordering that could be interpreted as a model of protein secondary structure. Hetero-polymers with the sequences of hydrophilic and hydrophobic residues characteristic for α- and β-type proteins undergo a somewhat sharper (though continuous) collapse transition to a dense globular state with elements of local ordering controlled by the sequence. The helical pattern induces more secondary structure than the β-type pattern. For all examined sequences the level of local ordering was lower than the average secondary structure content of globular proteins. The results are compared with other theoretical work and with known experimental facts. The implications for the reduced modeling of protein systems are briefly discussed. © 1997 John Wiley & Sons, Inc. Biopoly 42: 537–548, 1997  相似文献   

5.
The islet amyloid polypeptide (IAPP) and insulin are coproduced by the β-cells of the pancreatic islets of Langerhans. Both peptides can interact with negatively charged lipid membranes. The positively charged islet amyloid polypeptide partially inserts into these membranes and subsequently forms amyloid fibrils. The amyloid fibril formation of insulin is also accelerated by the presence of negatively charged lipids, although insulin has a negative net charge at neutral pH-values. We used water-polymer model interfaces to differentiate between the hydrophobic and electrostatic interactions that can drive these peptides to adsorb at an interface. By applying neutron reflectometry, the scattering-length density profiles of IAPP and insulin, as adsorbed at three different water-polymer interfaces, were determined. The islet amyloid polypeptide most strongly adsorbed at a hydrophobic poly-(styrene) surface, whereas at a hydrophilic, negatively charged poly-(styrene sulfonate) interface, the degree of adsorption was reduced by 50%. Almost no IAPP adsorption was evident at this negatively charged interface when we added 100 mM NaCl. On the other hand, negatively charged insulin was most strongly attracted to a hydrophilic, negatively charged interface. Our results suggest that IAPP is strongly attracted to a hydrophobic surface, whereas the few positive charges of IAPP cannot warrant a permanent immobilization of IAPP at a hydrophilic, negatively charged surface at an ionic strength of 100 mM. Furthermore, the interfacial accumulation of insulin at a hydrophilic, negatively charged surface may represent a favorable precondition for nucleus formation and fibril formation.  相似文献   

6.
The process of protein folding is obviously driven by forces exerted on the atoms of the amino‐acid chain. These forces arise from interactions with other parts of the protein itself (direct forces), as well as from interactions with the solvent (solvent‐induced forces). We present a statistical–mechanical formalism that describes both these direct and indirect, solvent‐induced thermodynamic forces on groups of the protein. We focus on 2 kinds of protein groups, commonly referred to as hydrophobic and hydrophilic. Analysis of this result leads to the conclusion that the forces on hydrophilic groups are in general stronger than on hydrophobic groups. This is then tested and verified by a series of molecular dynamics simulations, examining both hydrophobic alkanes of different sizes and hydrophilic moieties represented by polar‐neutral hydroxyl groups. The magnitude of the force on assemblies of hydrophilic groups is dependent on their relative orientation: with 2 to 4 times larger forces on groups that are able to form one or more direct hydrogen bonds.  相似文献   

7.
Hot spot residues contribute dominantly to protein-protein interactions. Statistically, conserved residues correlate with hot spots, and their occurrence can distinguish between binding sites and the remainder of the protein surface. The hot spot and conservation analyses have been carried out on one side of the interface. Here, we show that both experimental hot spots and conserved residues tend to couple across two-chain interfaces. Intriguingly, the local packing density around both hot spots and conserved residues is higher than expected. We further observe a correlation between local packing density and experimental deltadeltaG. Favorable conserved pairs include Gly coupled with aromatics, charged and polar residues, as well as aromatic residue coupling. Remarkably, charged residue couples are underrepresented. Overall, protein-protein interactions appear to consist of regions of high and low packing density, with the hot spots organized in the former. The high local packing density in binding interfaces is reminiscent of protein cores.  相似文献   

8.
The solvent structure and behavior around a protein were examined by analyzing a trajectory of molecular dynamics simulation of thetrp-holorepressor in a periodic box of water. The calculated selfdiffusion coefficient indicated that the solvent within 10 Å of the protein had lower mobility. Examination of the solvent diffusion around different atoms of different kinds of residues showed no general tendency. Thisfact suggested that the solvent mobility is not influenced significantly bythe kind of the atom or residue they solvated. Distribution analysis aroundthe protein revealed two peaks of water oxygen: a sharp one at 2.8 Å around polar and charged atoms and a broad one at ~3.4 Å aroundapolar atoms. The former was stabilized by water–protein hydrogen bonds, and the latter was stabilized by water-lwater hydrogen bonds, suggesting the existence of a hydrophobic shell. An analysis of protein atom–water radial distribution functions confirmed these shell structures around polar or charged atoms and apolar ones. © 1993 Wiley-Liss, Inc.  相似文献   

9.
Understanding protein adsorption kinetics to surfaces is of importance for various environmental and biomedical applications. Adsorption of bovine serum albumin to various self-assembled monolayer surfaces including neutral and charged hydrophilic and hydrophobic surfaces was investigated using in-situ combinatorial quartz crystal microbalance with dissipation and spectroscopic ellipsometry. Adsorption of bovine serum albumin varied as a function of surface properties, bovine serum albumin concentration and pH value. Charged surfaces exhibited a greater quantity of bovine serum albumin adsorption, a larger bovine serum albumin layer thickness, and increased density of bovine serum albumin protein compared to neutral surfaces at neutral pH value. The quantity of adsorbed bovine serum albumin protein increased with increasing bovine serum albumin concentration. After equilibrium sorption was reached at pH 7.0, desorption of bovine serum albumin occurred when pH was lowered to 2.0, which is below the isoelectric point of bovine serum albumin. Our data provide further evidence that combinatorial quartz crystal microbalance with dissipation and spectroscopic ellipsometry is a sensitive analytical tool to evaluate attachment and detachment of adsorbed proteins in systems with environmental implications.  相似文献   

10.
We introduce a microscopic model of a lipid with a charged headgroup and flexible hydrophobic tails, a neutral solvent, and counter ions. Short-ranged interactions between hydrophilic and hydrophobic moieties are included as are the Coulomb interactions between charges. Further, we include a short-ranged interaction between charges and neutral solvent, which mimics the short-ranged, thermally averaged interaction between charges and water dipoles. We show that the model of the uncharged lipid displays the usual lyotropic phases as a function of the relative volume fraction of the headgroup. Choosing model parameters appropriate to dioleoylphosphatidylethanolamine in water, we obtain phase behavior that agrees well with experiment. Finally we choose a solvent concentration and temperature at which the uncharged lipid exhibits an inverted hexagonal phase and turn on the headgroup charge. The lipid system makes a transition from the inverted hexagonal to the lamellar phase, which is related to the increased waters of hydration correlated with the increased headgroup charge via the charge-solvent interaction. The polymorphism displayed upon variation of pH mimics that of the behavior of phosphatidylserine.  相似文献   

11.
A two-dimensional off-lattice protein model with two species of monomers, hydrophobic and hydrophilic, was studied. Low-energy configurations in the model were optimized using the improved energy landscape paving (ELP+) method. In ELP+, the energy landscape paving (ELP) was first applied to search for the low-energy states. After the ELP led to the basins of the local energy minima, the additional degree-of-freedom of bond length was introduced, and the gradient descent method was then used to search for lower energy states near the local minima. Numerical results show that the proposed methods are quite effective for finding the ground states of proteins. A comparison between ELP+ and other methods is made.  相似文献   

12.
The native three-dimensional structure of a single protein is determined by the physicochemical nature of its constituent amino acids. The 20 different types of amino acids, depending on their physicochemical properties, can be grouped into three major classes: hydrophobic, hydrophilic, and charged. The anatomy of the weighted and unweighted networks of hydrophobic, hydrophilic, and charged residues separately for a large number of proteins were studied. Results showed that the average degree of the hydrophobic networks has a significantly larger value than that of hydrophilic and charged networks. The average degree of the hydrophilic networks is slightly higher than that of the charged networks. The average strength of the nodes of hydrophobic networks is nearly equal to that of the charged network, whereas that of hydrophilic networks has a smaller value than that of hydrophobic and charged networks. The average strength for each of the three types of networks varies with its degree. The average strength of a node in a charged network increases more sharply than that of the hydrophobic and hydrophilic networks. Each of the three types of networks exhibits the "small-world" property. Our results further indicate that the all-amino-acids networks and hydrophobic networks are of assortative type. Although most of the hydrophilic and charged networks are of the assortative type, few others have the characteristics of disassortative mixing of the nodes. We have further observed that all-amino-acids networks and hydrophobic networks bear the signature of hierarchy, whereas the hydrophilic and charged networks do not have any hierarchical signature.  相似文献   

13.
The probability of occurrence of helix and β-sheet residues in 47 globular proteins was determined as a function of local hydrophobicity, which was defined by the sum of the Nozaki-Tanford transfer free energies at two nearest-neighbors on both sides of the amino acid sequence. In general, hydrophilic amino acids favor neither helix nor β-sheet formations when neighbor residues are also hydrophilic but favor helix formation at higher local hydrophobicity. On the other hand, some hydrophobic amino acids such as Met, Leu, and Ile favor helix formation when neighbor residues are hydrophilic. None of the hydrophobic amino acids favor β-sheet formation with hydrophilic neighbors, but most of them strongly favor β-sheet formation at high local hydrophobicity. When the average of 20 amino acids is taken, both helix and β-sheet residue probabilities are higher at higher local hydrophobicity, although the increase is steeper for β-sheets. Therefore, β-sheet formation is more influenced by local hydrophobicity than helix formation. Generally, helices are nearer the surface and tend to have hydrophilic and hydrophobic faces at opposite sides. The tendency of alternating regions of hydrophilic and hydrophobic residues in a helical sequence was revealed by calculating the correlation of the Nozaki-Tanford values. Such amphipathic helices may be important in protein–protein and protein–lipid interactions and in forming hydrophilic channels in the membrane. The choice of 30 nonhomologous proteins as the data set did not alter the above results.  相似文献   

14.
The structure and dynamics of the fatty acid binding cavity in I-FABP (rat intestinal fatty acid binding protein) were analyzed. In the crystal structure of apo I-FABP, the probe occupied cavity volume and surface are 539+/-8 A3 and 428 A2, respectively (1.4 A probe). A total of 31 residues contact the cavity with their side chains. The side-chain cavity surface is partitioned according to the residue type as follows: 36-39% hydrophobic, 21-25% hydrophilic, and 37-43% neutral or ambivalent. Thus, the cavity surface is neither like a typical protein interior core, nor is like a typical protein external surface. All hydrophilic residues that contact the cavity-with the exception of Asp74-are clustered on the one side of the cavity. The cavity appears to expand its hydrophobic surface upon fatty acid binding on the side opposite to this hydrophilic patch. In holo I-FABP the fatty acid chain interactions with the hydrophilic side chains are mediated by water molecules. Molecular dynamics (MD) simulation of fully solvated apo I-FABP showed global conformational changes of I-FABP, which resulted in a large, but seemingly transient, exposure of the cavity to the external solvent. The packing density of the side chains lining the cavity, studied by Voronoi volumes, showed the presence of two distinctive small hydrophobic cores. The MD simulation predicts significant structural perturbations of the cavity on the subnanosecond time scale, which are capable of facilitating exchange of I-FABP internal water.  相似文献   

15.
Protein folding is directed by the sequence of sidechains along the polypeptide backbone, but despite this the developement of sidechain interactions during folding is not well understood. Here, the thiol-active reagent, dithio-nitrobenzoic acid (DTNB), is used to probe the exposure of the cysteine sidechain thiols in the kinetic folding intermediates of the N-terminal domain of phosphoglycerate kinase (N-PGK) and a number of conservative (I-, L-, or V-to-C) single cysteine variants. Rapid dilution of chemically denatured protein into folding conditions in the presence of DTNB allowed the degree of sidechain protection in any rapidly formed intermediate to be determined through the analysis of the kinetics of labelling. The protection factors derived for the intermediate(s) were generally small (<25), indicating only partial burial of the sidechains. The distribution of protection parallels the previously reported backbone amide protection for the folding intermediate of N-PGK. These observations are consistent with the hypothesis that such intermediates resemble molten globule states; i.e. with native-like backbone hydrogen bonding and overall tertiary structure, but with the sidechains that make up the hydrophobic protein core dynamic and intermittently solvent exposed. The success of the competition technique in characterizing this kinetic intermediate invites application to other model systems.  相似文献   

16.
In our previous study, we have shown that the microenvironments around conserved amino acids are also conserved in protein families (Bandyopadhyay and Mehler, Proteins 2008; 72:646–659). In this study, we have hypothesized that amino acids perform similar functions when embedded in a certain type of protein microenvironment. We have tested this hypothesis on the microenvironments around disulfide‐bridged cysteines from high‐resolution protein crystal structures. Although such cystines mainly play structural role in proteins, in certain enzymes they participate in catalysis and redox reactions. We have performed and report a functional annotation of enzymatically active cystines to their respective microenvironments. Three protein microenvironment clusters were identified: (i) buried‐hydrophobic, (ii) exposed‐hydrophilic, and (iii) buried‐hydrophilic. The buried‐hydrophobic cluster encompasses a small group of 22 redox‐active cystines, mostly in alpha‐helical conformations in a –C‐x‐x‐C‐ motif from the Oxido‐reductase enzyme class. All these cystines have high strain energy and near identical microenvironments. Most of the active cystines in hydrolase enzyme class belong to buried hydrophilic microenvironment cluster. In total there are 34 half‐cystines detected in buried hydrophilic cluster from hydrolases, as a part of enzyme active site. Even within the buried hydrophilic cluster, there is clear separation of active half‐cystines between surface exposed part of the protein and protein interior. Half‐cystines toward the surface exposed region are higher in number compared to those in protein interior. Apart from cystines at the active sites of the enzymes, many more half‐cystines were detected in buried hydrophilic cluster those are part of the microenvironment of enzyme active sites. However, no active half‐cystines were detected in extremely hydrophilic microenvironment cluster, that is, exposed hydrophilic cluster, indicating that total exposure of cystine toward the solvent is not favored for enzymatic reactions. Although half‐cystines in exposed‐hydrophilic clusters occasionally stabilize enzyme active sites, as a part of their microenvironments. Analysis performed in this work revealed that cystines as a part of active sites in specific enzyme families or folds share very similar protein microenvironment regions, despite of their dissimilarity in protein sequences and position specific sequence conservations. Proteins 2016; 84:1576–1589. © 2016 Wiley Periodicals, Inc.  相似文献   

17.
Nanda V  Schmiedekamp A 《Proteins》2008,70(2):489-497
Proteins fold and maintain structure through the collective contributions of a large number of weak, noncovalent interactions. The hydrogen bond is one important category of forces that acts on very short distances. As our knowledge of protein structure continues to expand, we are beginning to appreciate the role that weak carbon-donor hydrogen bonds play in structure and function. One property that differentiates hydrogen bonds from other packing forces is propensity for forming a linear donor-hydrogen-acceptor orientation. To ascertain if carbon-donor hydrogen bonds are able to direct acceptor linearity, we surveyed the geometry of interactions specifically involving aromatic sidechain ring carbons in a data set of high resolution protein structures. We found that while donor-acceptor distances for most carbon donor hydrogen bonds were tighter than expected for van der Waals packing, only the carbons of histidine showed a significant bias for linear geometry. By categorizing histidines in the data set into charged and neutral sidechains, we found only the charged subset of histidines participated in linear interactions. B3LYP/6-31G**++ level optimizations of imidazole and indole-water interactions at various fixed angles demonstrates a clear orientation dependence of hydrogen bonding capacity for both charged and neutral sidechains. We suggest that while all aromatic carbons can participate in hydrogen bonding, only charged histidines are able to overcome protein packing forces and enforce linear interactions. The implications for protein modeling and design are discussed.  相似文献   

18.
A series of novel amphiphilic macromolecules composed of alkyl chains as the hydrophobic block and poly(ethylene glycol) as the hydrophilic block were designed to inhibit highly oxidized low density lipoprotein (hoxLDL) uptake by synthesizing macromolecules with negatively charged moieties (ie, carboxylic acids) located in the two different blocks. The macromolecules have molecular weights around 5,500 g/mol, form micelles in aqueous solution with an average size of 20-35 nm, and display critical micelle concentration values as low as 10(-7) M. Their charge densities and hydrodynamic size in physiological buffer solutions correlated with the hydrophobic/ hydrophilic block location and quantity of the carboxylate groups. Generally, carboxylate groups located in the hydrophobic block destabilize micelle formation more than carboxylate groups in the hydrophilic block. Although all amphiphilic macromolecules inhibited unregulated uptake of hoxLDL by macrophages, inhibition efficiency was influenced by the quantity and location of the negatively charged-carboxylate on the macromolecules. Notably, negative charge is not the sole factor in reducing hoxLDL uptake. The combination of smaller size, micellar stability and charge density is critical for inhibiting hoxLDL uptake by macrophages.  相似文献   

19.
The sequence of events in the refolding pathway of barnase has been analysed to search for general principles in protein folding. There appears to be a correlation between burying hydrophobic surface area and early folding events. All the regions that fold early interact extensively with the beta-sheet. These interactions involve predominantly hydrophobic interactions and the burial of very extensive hydrophobic areas in which multiple, close, hydrophobic-hydrophobic contacts are established around a central group of aliphatic residues. There is no burial of hydrophilic residues in these regions; those that are partly screened from the solvent make hydrogen bonds. All the regions or interactions that are made late in the folding pathway do not make extensive contacts with the beta-sheet. Their buried hydrophobic regions lack a central hydrophobic residue or residues around which other hydrophobic residues pack. Further, in some of these regions there is an extensive burial of hydrophilic residues. The results are consistent with one of the earlier events in protein folding being the local formation of native-like secondary structure elements driven by local hydrophobic surface burial. A possible candidate for an initiation site is a beta-hairpin between beta-strands 3 and 4 that is conserved in the microbial ribonuclease family. A comparison of structures in this family shows that those regions that can be superimposed, or have sequence homology, correspond to elements of structure that are formed and interact with each other early in the folding pathway, suggesting that some of these residues could be involved in directing the folding process. The data on barnase combined with results from other laboratories suggest the following tentative conclusions for the refolding of small monomeric proteins. (1) The refolding pathway is, at least in part, sequential and of compulsory order. (2) Secondary structure formation is driven by local hydrophobic surface burial and precedes the formation of most tertiary interactions. These elements are then stabilized and sometimes elongated by tertiary interactions. It is plausible that there are stop signals encoded in the linear sequence that prevent the elongation of isolated secondary structure elements in solution to a larger extent than is found in the folded protein. (3) Many tertiary interactions are not very constrained in the intermediate but become more and more defined as the hydrophobic cores consolidate, loop structures form and the configuration of surface residues takes place. The interactions between different elements of secondary structure are the last ones to be consolidated while the interactions within the secondary structure elements are consolidated earlier.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

20.
Cytochrome b5 is a membrane protein that comprises two fragments: one is water-soluble and heme-containing, and the other is hydrophobic and membrane-embedded. The function of electron transfer is performed by the former whose crystal structure is known; however, its conformational states when in the membrane field and interacting with other proteins are still to be studied. Previously, we proposed water-alcohol mixtures for modeling the effect of membrane surface on proteins, and used this approach to study the conformational behavior of positively charged cytochrome c as well as relatively neutral retinol-binding protein also functioning in the field of negatively charged membrane. The current study describes the conformational behavior of the negatively charged water-soluble fragment of cytochrome b5 as dependent on pH. Decreasing pH was shown to transform the fragment state from native to intermediate, similar to the molten globule reported earlier for other proteins in aqueous solutions: at pH 3.0, the fragment preserved a pronounced secondary structure and compactness but lost its rigid tertiary structure. A possible role of this intermediate state in cytochrome b5 functioning is discussed.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号