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 421 毫秒
1.
The structure of the sulphydryl protease, actinidin, after refinement at 1.7 Å resolution, is described. The positions of most of the 1666 atoms have been determined with an accuracy better than 0.1 Å; only two residues (219 and 220) and the side-chain of a third (87) cannot be seen. In addition, the model contains 272 solvent molecules, all taken as water, except one which may be an ammonium ion. Atomic B values give a good indication of the mobility of different parts of the structure. Actinidin has a double domain structure, with one domain mostly helical in its secondary structure, and the other domain built around a twisted β-sheet. The geometry of hydrogen bonds in helices, β-structure and turns has been analysed. All are significantly non-linear, with the angle N-?…O ~160 °. Carbonyl groups are tilted outwards from the axis of each helix, the tilting apparently unaffected by whether or not additional hydrogen bonds are made (e.g. to water or side-chain atoms). Each domain is folded round a substantial core of non-polar side-chains, but the interface between domains is mostly polar. Interactions across this interface involve a network of eight buried water molecules, the buried carboxyl and amino groups of Glu35, Glu50, Lys181 and Lys17, other polar side-chains and a few hydrophobic groups. One other internal charged side-chain, that of Glu52, is adjacent to a buried solvent molecule, probably an ammonium ion. Other side-chain environments are described. One proline residue has a cis configuration. The sulphydryl group is oxidized, probably to SO2?, with one oxygen atom clearly visible but the other somewhat less certain. The active site geometry is otherwise compatible with the mechanism proposed by Drenth et al. (1975,1976) for papain. The positions of the 272 solvent molecules are described. The best-ordered water molecules are those that are internal (total of 17), in surface pockets, or in the intermolecular contact regions. These generally form three or four hydrogen bonds, two to proton acceptors and one or two to proton donors. Other water molecules make water bridges on the surface, sometimes covering the exposed edges of non-polar groups. Intermolecular contacts involve few protein atoms, but many water molecules.  相似文献   

2.
The networks of solvent hydrogen bonds around polar and apolar amino acids have been studied by computer simulation techniques using a non-pair additive model for the water molecules interactions. Analysis of the simulated aqueous solutions has shown the presence of water molecules which (a) form a bridge around individual polar solute atoms (self-bridging loops) and (b) form chains between different polar solute atoms (polar bridging chains). Some of these networks associated with polar solute atoms from pentagons but 4, 6 and 7 sided polygons are also seen. The water molecule close to apolar solute atoms (<4.0 Å) also form irregular networks with polygons of 4, 5, 6 and 7 sides. These networks are compared with those found experimentally in ice, clathrates and crystal hydrates of macromolecules.  相似文献   

3.
In a selected set of 44 high-resolution, non-homologous protein structures, the intramolecular hydrogen bonds or salt bridges formed by ionizable amino acid side chains were identified and analyzed. The analysis was based on the investigation of several properties of the involved residues such as their solvent exposure, their belonging to a certain secondary structural element, and their position relative to the N- and C-termini of their respective structural element. It was observed that two-thirds of the interactions made by basic or acidic side chains are hydrogen bonds to polar uncharged groups. In particular, the majority (78%) of the hydrogen bonds between ionizable side chains and main chain polar groups (sch:mch bonds) involved at least one buried atom, and in 42% of the cases both interacting atoms were buried. In α-helices, the sch:mch bonds observed in the proximity of the C- and N-termini show a clear preference for acidic and basic side chains, respectively. This appears to be due to the partial charges of peptide group atoms at the termini of α-helices, which establish energetically favorable electrostatic interactions with side chain carrying opposite charge, at distances even greater than 4.5 Å. The sch:mch interactions involving ionizable side chains that belong either to β-strands or to the central part of α-helices are based almost exclusively on basic residues. This results from the presence of main chain carbonyl oxygen atoms in the protein core which have unsatisfied hydrogen bonding capabilities.  相似文献   

4.
The structure of alpha-lytic protease, a serine protease produced by the bacterium Lysobacter enzymogenes, has been refined at 1.7 A resolution. The conventional R-factor is 0.131 for the 14,996 reflections between 8 and 1.7 A resolution with I greater than or equal to 2 sigma (I). The model consists of 1391 protein atoms, two sulfate ions and 156 water molecules. The overall root-meansquare error is estimated to be about 0.14 A. The refined structure was compared with homologous enzymes alpha-chymotrypsin and Streptomyces griseus protease A and B. A new sequence numbering was derived based on the alignment of these structures. The comparison showed that the greatest structural homology is around the active site residues Asp102, His57 and Ser195, and that basic folding pathways are maintained despite chemical changes in the hydrophobic cores. The hydrogen bonds in the structure were tabulated and the distances and angles of interaction are similar to those found in small molecules. The analysis also revealed the presence of close intraresidue interactions. There are only a few direct intermolecular hydrogen bonds. Most intermolecular interactions involve bridging solvent molecules. The structural importance of hydrogen bonds involving the side-chain of Asx residues is discussed. All the negatively charged groups have a counterion nearby, while the excess positively charged groups are exposed to the solvent. One of the sulfate ions is located near the active site, whereas the other is close to the N terminus. Of the 156 water molecules, only seven are not involved in a hydrogen bond. Six of these have polar groups nearby, while the remaining one is in very weak density. There are nine internal water molecules, consisting of two monomers, two dimers and one trimer. No significant second shell of solvent is observed.  相似文献   

5.
Luise A  Falconi M  Desideri A 《Proteins》2000,39(1):56-67
A system containing the globular protein azurin and 3,658 water molecules has been simulated to investigate the influence on water dynamics exerted by a protein surface. Evaluation of water mean residence time for elements having different secondary structure did not show any correlation. Identically, comparison of solvent residence time for atoms having different charge and polarity did not show any clear trend. The main factor influencing water residence time in proximity to a specific site was found to be its solvent accessibility. In detail for atoms belonging to lateral chains and having solvent-accessible surface lower than approximately 16 A(2)a relation is found for which charged and polar atoms are surrounded by water molecules characterized by residence times longer than the non polar ones. The involvement of the low accessible protein atom in an intraprotein hydrogen bond further modulates the length of the water residence time. On the other hand for surfaces having high solvent accessibility, all atoms, independently of their character, are surrounded by water molecules which rapidly exchange with the bulk solvent. Proteins 2000;39:56-67.  相似文献   

6.
The structure of human erythrocytic carbonic anhydrase II has been refined by constrained and restrained structure–factor least-squares refinement at 2.0 Å resolution. The conventional crystallographic R value is 17.3%. Of 167 solvent molecules associated with the protein, four are buried and stabilize secondary structure elements. The zinc ion is ligated to three histidyl residues and one water molecule in a nearly tetrahedral geometry. In addition to the zinc-bound water, seven more water molecules are identified in the active site. Assuming that Glu-106 is deprotonated at pH 8.5, some of the hydrogen bond donor–acceptor relations in the active site can be assigned and are described here in detail. The Oγ1 atom of Thr-199 donates its proton to the Oε1 atom of Glu-106 and can function as a hydrogen bond acceptor only in additional hydrogen bonds.  相似文献   

7.
Contribution of electrostatic interactions to stability of BPTI orthorhombic, pig-insulin cubic crystals, and horse L ferritin crystals was evaluated with numerical calculation of Poisson-Boltzmann equation based on a dielectric model. The stability of a ferritin molecule (24-mer) composed of 24 subunits was also evaluated. It was found that the surface charge-charge interactions at separation distances (< 5 Å) were insensitive to variations in the ionic strength, and thus stabilized assembled states of the proteins (i.e., crystalline state and oligomeric state). It was also revealed that the charge density and the packing of the protein crystals were largely responsible for the ionic strength dependence of the crystal stability. The stability of the 5PTI crystalline state with a high charge density drastically increased as the concentration of the solvent ions increased. In contrast, that of the insulin crystal with a low charge density and large solvent region was insensitive to changes in the ionic concentration. The electrostatic interaction between ferritin 24-mers was attributed to two salt bridges mediated by Cd ion. For the stability of the ferritin 24-mer, which is evolutionally designed, the electrostatic stabilization between the subunits was attributed to polar bonds such as buried salt bridges or hydrogen bonds, which occasionally yielded more than 5 kcal/mol and were numerous and very strong compared with the bonds between molecules in the 5PTI and 9INS crystals.By analyzing the atomic charge-charge interactions in detail, it was found that charge pairs separated by less than 3 Å, such as hydrogen bonds, dominantly stabilize the assembled states, and that pairs 3 to 5 Å apart were also important. The stability of the assembled states evaluated by the total EET was determined by the fine balance between the two competing contributions arising from the stabilizing atoms and the destabilizing atoms.Changes of the ASA and hydration free energy were also evaluated in accordance with the process of the subunit assembly. The change of hydration free energy, which was very large (i.e., ~+ 100 kcal/mol/subunit) and unfavorable for the assembly, was proportional to the electrostatic hydration energy (i.e., Born energy change in the hydration process). Hydrophobic groups were likely to appear more frequently than hydrophilic groups at the interfaces.This study offers a method which can improve the stability of protein crystals by introducing polar or charged residues that are properly designed to form specific hydrogen bonds or salt bridges between neighboring protein molecules. This method is also applicable to crystallography, because it improves refinement of protein structures in crystals by taking the inter-protein interactions into account.  相似文献   

8.
Park S  Saven JG 《Proteins》2005,60(3):450-463
Buried solvent molecules are common in the core of globular proteins and contribute to structural stability. Folding necessitates the burial of polar backbone atoms in the protein core, whose hydrogen-bonding capacities should be satisfied on average. Whereas the residues in alpha-helices and beta-sheets form systematic main-chain hydrogen bonds, the residues in turns, coils and loops often contain polar atoms that fail to form intramolecular hydrogen bonds. The statistical analysis of 842 high resolution protein structures shows that well-resolved, internal water molecules preferentially reside near residues without alpha-helical and beta-sheet secondary structures. These buried waters most often form primary hydrogen bonds to main-chain atoms not involved in intramolecular hydrogen bonds, providing strong evidence that hydrating main-chain atoms is a key structural role of buried water molecules. Additionally, the average B-factor of protein atoms hydrogen-bonded to waters is smaller than that of protein atoms forming intramolecular hydrogen bonds, and the average B-factor of water molecules involved in primary hydrogen bonds with main-chain atoms is smaller than the average B-factor of water molecules involved in secondary hydrogen bonds to protein atoms that form concurrent intramolecular hydrogen bonds. To study the structural coupling between internal waters and buried polar atoms in detail we simulated the dynamics of wild-type FKBP12, in which a buried water, Wat137, forms one side-chain and multiple main-chain hydrogen bonds. We mutated E60, whose side-chain hydrogen bonds with Wat137, to Q, N, S or A, to modulate the multiplicity and geometry of hydrogen bonds to the water. Mutating E60 to a residue that is unable to form a hydrogen bond with Wat137 results in reorientation of the water molecule and leads to a structural readjustment of residues that are both near and distant to the water. We predict that the E60A mutation will result in a significantly reduced affinity of FKBP12 for its ligand FK506. The propensity of internal waters to hydrogen bond to buried polar atoms suggests that ordered water molecules may constitute fundamental structural components of proteins, particularly in regions where alpha-helical or beta-sheet secondary structure is not present.  相似文献   

9.

Background  

The hydrogen bond patterns between mainchain atoms in protein structures not only give rise to regular secondary structures but also satisfy mainchain hydrogen bond potential. However, not all mainchain atoms can be satisfied through hydrogen bond interactions that arise in regular secondary structures; in some locations sidechain-to-mainchain hydrogen bonds are required to provide polar group satisfaction. Buried polar residues that are hydrogen-bonded to mainchain amide atoms tend to be highly conserved within protein families, confirming that mainchain architecture is a critical restraint on the evolution of proteins. We have investigated the stabilizing roles of buried polar sidechains on the backbones of protein structures by performing an analysis of solvent inaccessible residues that are entirely conserved within protein families and superfamilies and hydrogen bonded to an equivalent mainchain atom in each family member.  相似文献   

10.
The structure of horse methaemoglobin at 2-0 A resolution   总被引:13,自引:0,他引:13  
The structure of horse methaemoglobin has been redetermined by phase extension and refinement. This has improved our knowledge of the haem geometry and the stereochemistry of the interfaces between the subunits, and confirmed the disorder of the C-terminal residues. Using new four-circle diffractometer data between the limiting spheres of radius 10 and 2.0 Å?1, the co-ordinates determined by Perutz et al. (1968a,b) were subjected to successive cycles of real-space refinement into electron density maps calculated with observed ¦F¦ values and phases derived from the latest refined model, until the reliability index had dropped from an initial value of 0.45 to 0.23. The positions of the iron atoms relative to the planes of the porphyrin rings were refined separately, and checked by Fourier syntheses based on anomalous scattering and by difference Fourier syntheses calculated with coefficients from which the iron contributions had been removed. The general root-mean-squared error in atomic positions is 0.32 Å; the probable error in the displacement of the iron atoms from the porphyrin planes is 0.06 Å. The difference Fourier synthesis, obtained after refinement of the protein was complete, showed 41 bound water molecules per asymmetric unit and also revealed five errors in amino acid sequence, one of which was confirmed chemically.The secondary structures of the subunits are stabilized by hydrogen bonds formed by main-chain NH and CO groups either with each other or with nearby polar side-chains. There are few internal hydrogen bonds linking the various chain segments; many of the external polar side-chains help to stabilize the tertiary structure by forming hydrogen bonds with each other or through bound water molecules. Several of the helical segments are irregular and the terminal residues are disordered. The contacts between the subunits are more polar than the earlier 2.8 Å map had led us to believe, because it had failed to show up the 15 bound water molecules at the α1β1 and the four at the α1β2 contact. Their inclusion has raised the number of hydrogen bonds between neighbouring subunits at α1β1 from five to 17 or possibly 19, and at α1β2 from two to six or possibly seven. The remaining 22 water molecules are distributed over the internal cavity and the molecular surface; most of them make hydrogen bonds with at least two polar groups of the protein. Despite several amino acid differences, the structure of the α1β1 contact, including the bound water, is the same as in human deoxyhaemoglobin (Fermi, 1975).  相似文献   

11.
Structure and refinement of penicillopepsin at 1.8 A resolution   总被引:15,自引:0,他引:15  
Penicillopepsin, the aspartyl protease from the mould Penicillium janthinellum, has had its molecular structure refined by a restrained-parameter least-squares procedure at 1.8 Å resolution to a conventional R-factor of 0.136. The estimated co-ordinate accuracy for the majority of the 2363 atoms of the enzyme is better than 0.12 Å. The average atomic thermal vibration parameter, B, for the atoms of the enzyme is 14.5 Å2. One determining factor of this low average B value is the large central hydrophobic core, in which there are two prominent clusters of aromatic residues, one of nine, the other of seven residues. The N and C-terminal domains of penicillopepsin display an approximate 2-fold symmetry: 70 residue pairs are topologically equivalent, related by a rotation of 177 ° and a translation of 1.2 Å. The analysis of the secondary structural features of the molecule reveals non-linear hydrogen bonding. In penicillopepsin, there is no difference in the mean hydrogen-bond parameters for the elements of α-helix, parallel or antiparallel β-pleated sheet. The mean values for these structural elements are: NO, 2.90 Å; NHO, 1.95 Å; N?O, 160 °. The average hydrogen-bond parameters of the reverse β-turns and the 310 helices are distinctly different from the above values. The analysis of sidechain conformational angles χ1 and χ2 penicillopepsin and other enzyme structures refined in this laboratory shows much narrower distributions as compared with those compiled from unrefined protein structures. The close proximity of the carboxyl groups of Asp33 and Asp213 suggests that they share a proton in a tight hydrogen-bonded environment (Asp33OD2 to Asp213OD1 is 2.87 Å). There are several solvent molecules in the active site region and, in particular, O39 forms hydrogen-bonded interactions with both aspartate residues. The disposition of the two carboxyl groups suggests that neither is likely to be involved in a direct nucleophilic attack on the scissile bond of a substrate. The average atomic B-factors of the residues in this region of the molecule are between 5 and 8 Å2, confirming the proposal that conformational mobility of the active site residues has no role in the enzymatic mechanism. However, conformational mobility of neighbouring regions of the molecule e.g. the “flap” containing Tyr75, is verified by the high B-factors for those residues. The positions of 319 solvent sites per asymmetric unit have been selected from difference electron density maps and refined. Thirteen have been classified as internal, and several of these may have key roles during catalysis. The positively charged Nζ atom of Lys304 forms hydrogen bonds to the carboxylate of Asp14 (internal ion pair) and to two internal water molecules O5 and O25. The protonated side-chain of Asp300 forms a hydrogen bond to Thr214O, 2.78 Å, and is the recipient of a hydrogen bond from a surface pocket water molecule O46. There is no possibility for direct interaction between Asp300 and Lys304 without large conformational changes of their environment. The intermolecular packing involves many protein-protein contacts (66 residues) with a large number of solvent molecules involved in bridging between polar residues at the contact surface. The penicillopepsin molecules resemble an approximate hexagonal close-packing of spheres with each molecule having 12 “nearest” neighbours.  相似文献   

12.
The experimental determination of protein compressibility reflects both the protein intrinsic compressibility and the difference between the compressibility of water in the protein hydration shell and bulk water. We use molecular dynamics simulations to explore the dependence of the isothermal compressibility of the hydration shell surrounding globular proteins on differential contributions from charged, polar, and apolar protein-water interfaces. The compressibility of water in the protein hydration shell is accounted for by a linear combination of contributions from charged, polar, and apolar solvent-accessible surfaces. The results provide a formula for the deconvolution of experimental data into intrinsic and hydration contributions when a protein of known structure is investigated. The physical basis for the model is the variation in water density shown by the surface-specific radial distribution functions of water molecules around globular proteins. The compressibility of water hydrating charged atoms is lower than bulk water compressibility, the compressibility of water hydrating apolar atoms is somewhat larger than bulk water compressibility, and the compressibility of water around polar atoms is about the same as the compressibility of bulk water. We also assess whether hydration water compressibility determined from small compound data can be used to estimate the compressibility of hydration water surrounding proteins. The results, based on an analysis from four dipeptide solutions, indicate that small compound data cannot be used directly to estimate the compressibility of hydration water surrounding proteins.  相似文献   

13.
Molecular dynamics simulations of Clostridium pasteurianum rubredoxin in the oxidized and reduced forms have been performed. Good agreement between both forms and crystal data has been obtained (rms deviation of backbone atoms of 1.06 and 1.42 Å, respectively), which was due in part to the use of explicit solvent and counterions. The reduced form exhibits an unexpected structural change: the redox site becomes much more solvent-accessible, so that water enters a channel between the surface and the site, but with little actual structural rearrangement (the rms deviation of backbone atoms between the oxidized and reduced is 0.77 Å). The increase in solvent accessibility is also seen, although to a much lesser extent, between the oxidized and reduced crystal structures of Pyrococcus furiosus rubredoxin, but no high resolution crystal or nuclear magnetic resonance solution data exist for reduced C. pasteurianum rubredoxin. The electrostatic potential at the iron site and fluctuations in the potential, which contribute to both the redox and electron transfer properties, have also been evaluated for both the oxidized and the reduced simulations. These results show that the backbone plays a significant role (62–70 kcall/mol/e) and the polar sidechains contribute relatively little (0–4 kcal/mol/e) to the absolute electrostatic potential at the iron of rubredoxin for both forms. However, both groups contribute significantly to the change in redox state by becoming more polarized and more densely packed around the redox site upon reduction. Furthermore, these results show that the solvent becomes much more polarized in the reduced form than in the oxidized form, even excluding the penetrating water. Finally, the simulation indicates that the contribution of the charged side chains to the electrostatic potential is largely canceled by that of the counterions. © 1995 Wiley-Liss, Inc.  相似文献   

14.
Molecular interaction between dioxane and methanol involves certain polar and nonpolar bonding to form a one to one complex. Interatomic distances between hydrogen and oxygen within 3 Å have been considered as hydrogen bonding. Optimizations of the structures of dioxane-methanol complexes were carried out considering any spatial orientation of a methanol molecule around a chair/boat/twisted-boat conformation of dioxane. From 45 different orientations of dioxane and water, 23 different structures with different local minima were obtained and the structural characteristics like interatomic distances, bond angles, dihedral angles, dipole moment of each complex were discussed. The most stable structure, i.e., with minimum heat of formation is found to have a chair form dioxane, one O-H…O, and two C-H…O hydrogen bonds. In general, the O-H…O hydrogen bonds have an average distance of 1.8 Å while C-H…O bonds have 2.6 Å. The binding energy of the dioxane-methanol complex is found to be a linear function of number of O-H…O and C-H…O bonds, and hydrogen bond length. Graphical Abstract
Sixteen orientations of methanol around dioxane converge to six local minima including the global minima with one H-O…H and two C-H…O hydrogen bonds.  相似文献   

15.
Hydrogen bond interactions were surveyed in a set of protein structures. Compared to surface positions, polar side-chains at core positions form a greater number of intra-molecular hydrogen bonds. Furthermore, the majority of polar side-chains at core positions form at least one hydrogen bond to main-chain atoms that are not involved in hydrogen bonds to other main-chain atoms. Based on this structural survey, hydrogen bond rules were generated for each polar amino acid for use in protein core design. In the context of protein core design, these prudent polar rules were used to eliminate from consideration polar amino acid rotamers that do not form a minimum number of hydrogen bonds. As an initial test, the core of Escherichia coli thioredoxin was selected as a design target. For this target, the prudent polar strategy resulted in a minor increase in computational complexity compared to a strategy that did not allow polar residues. Dead-end elimination was used to identify global minimum energy conformations for the prudent polar and no polar strategies. The prudent polar strategy identified a protein sequence that was thermodynamically stabilized by 2.5 kcal/mol relative to wild-type thioredoxin and 2.2 kcal/mol relative to a thioredoxin variant whose core was designed without polar residues.  相似文献   

16.
Molecular dynamics (MD) simulations were used to characterize the non-cooperative denaturation of the molten globule A-state of human alpha-lactalbumin by urea. A solvent of explicit urea and water molecules was used, corresponding to a urea concentration of approximately 6M. Three simulations were performed at temperatures of 293K, 360K and 400K, with lengths of 2 ns, 8 ns and 8 ns respectively. The results of the simulations were compared with experimental data from NMR studies of human alpha-lactalbumin and related peptides. During the simulations, hydrogen bonds were formed from the protein to both urea and water molecules as intra-protein hydrogen bonds were lost. Urea was shown to compete efficiently with water as both a hydrogen bond donor and acceptor. Radial distribution functions of water and urea around hydrophobic side chain atoms showed a significant increase in urea molecules in the solvation shell as the side chains became exposed during denaturation. A considerable portion of the native-like secondary structure persisted throughout the simulations. However, in the simulations at 360K and 400K, there were substantial changes in the packing of aromatic and other hydrophobic side chains in the protein, and many native contacts were lost. The results suggest that during the non-cooperative denaturation of the molten globule, secondary structure elements are stabilized by non-specific, non-native interactions.  相似文献   

17.
The structure of a crystalline cellulose I–ethylenediamine complex has been determined by x-ray diffraction methods as part of an investigation of cellulose–solvent interaction. The complex studied is that formed when native ramie fibers are swollen in ethylenediamine and then vacuum-dried. The unit cell is monoclinic with dimensions a = 12.87 Å, b = 9.52 Å, c = 10.35 Å, and γ = 118.8°, and it contains disaccharide segments of two chains, with one ethylenediamine per glucose residue. The refined model contains parallel cellulose chains that are linked by hydrogen-bonded ethylenediamine molecules. The chains along the b-axis are packed in register, leading to stacks of chains analogous to those in chitin. All the hydroxyl groups are satisfactorily hydrogen-bonded and each ethylenediamine forms four donor and two acceptor hydrogen bonds. From this work it can be seen that the interaction of cellulose I with ethylenediamine involves scission of the intermolecular hydrogen bonds followed by disruption of the stacks of quarter-staggered chains.  相似文献   

18.
Studies on the interactions between L ‐O‐ phosphoserine, as one of the simplest fragments of membrane components, and the Cinchona alkaloid cinchonine, in the crystalline state were performed. Cinchoninium L ‐O‐phosposerine salt dihydrate (PhSerCin) crystallizes in a monoclinic crystal system, space group P21, with unit cell parameters: a = 8.45400(10) Å, b = 7.17100(10) Å, c = 20.7760(4) Å, α = 90°, β = 98.7830(10)°, γ = 90°, Z = 2. The asymmetric unit consists of the cinchoninium cation linked by hydrogen bonds to a phosphoserine anion and two water molecules. Intermolecular hydrogen bonds connecting phosphoserine anions via water molecules form chains extended along the b axis. Two such chains symmetrically related by twofold screw axis create a “channel.” On both sides of this channel cinchonine cations are attached by hydrogen bonds in which the atoms N1, O12, and water molecules participate. This arrangement mimics the system of bilayer biological membrane. Chirality 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

19.
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   

20.
Abstract

As part of our on-going development of a method, based upon distance geometry calculations, for predicting the structures of proteins from the known structures of their homologues, we have predicted the structure of the 176 residue Flavodoxin from Escherichia coli. This prediction was based upon the crystal structures of the homologous Flavodoxins from Anacystis nidulans, Chondrus crispus, Desulfovibrio vulgaris and Clostridium beijerinckii, whose sequence identities with Escherichia coli were 44%, 33%, 23% and 16%, respectively. A total of 13,043 distance constraints among the alpha-carbons of the Escherichia coli structure were derived from the sequence alignments with the known structures, together with 8,893 distance constraints among backbone and sidechain atoms of adjacent residues, 978 between the alpha-carbons and selected atoms of the flavin mononucleotide cofactor, 116 constraints to enforce conserved hydrogen bonds, and 452 constraints on the torsion angles in conserved residues. An ensemble of ten random Escherichia coli structures was computed from these constraints, with an average root mean square coordinate deviation (RMSD) among the alpha carbons of 0.85 Ångstroms (excluding the first 1 and last 6 residues, which have no corresponding residues in any of the homologues and hence were unconstrained); the corresponding average heavy-atom RMSD was 1.60 Å.

Since the distance geometry calculations were performed without hydrogen atoms, protons were added to the resulting structures and these structures embedded in a 50 × 50 × 40 Å solvent box with periodic boundary conditions. They were then subjected to a 20 picosecond dynamical simulated annealing procedure, starting at 300 K and gradually reduced to 10K, in which all the distance and torsion angle constraints were maintained by means of harmonic restraint functions. This was followed up by 1000 iterations of unrestrained conjugate gradients minimization. The goal of this energy refinement procedure was not to drastically modify the structures in an attempt at a priori prediction, but merely to improve upon the predictions obtained from the geometric constraints, particularly with regard to their local backbone and sidechain conformations and their hydrogen bonds. The resulting structures differed from the respective starting structures by an average of 1.52 Å in their heavy atom RMSD's, while the average RMSD among the heavy atoms of residues 2-170 increased slightly to 1.66 Å. We hope these structures will be good enough to enable the phase problem to be solved for the crystallographic data that is now being collected on this protein.  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