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 31 毫秒
1.
Conformational constraints of amino acid side chains in alpha-helices   总被引:3,自引:0,他引:3  
L Piela  G Nemethy  H A Scheraga 《Biopolymers》1987,26(8):1273-1286
The conformational freedom of amino acid side chains is strongly reduced when the side chains occur on an α-helix. A quantitative evaluation of this freedom has been carried out by means of conformational energy computations for all naturally occurring amino acids and for α-aminobutyric acid when they are placed in the middle of a right-handed poly(L-alanine) α-helix. One of the three possible rotameric states for rotation around the Cα ? Cβ bond (viz. g+) is excluded completely on the helix because of steric hindrance, and the relative populations of the other two rotamers (t and g?) are altered because of steric interactions and the reduction of hydrogen-bonding possibilities. The computed tendencies of the changes in distributions of rotamers, on going from an ensemble of all backbone conformations to the α-helix, agree with the observed tendencies in proteins. Minimum-energy side-chain conformations in an α-helix have been tabulated for use in conformational energy computations on polypeptides.  相似文献   

2.
A modification of the α-helix, termed the ω-helix, has four residues in one turn of a helix. We searched the ω-helix in proteins by the HELFIT program which determines the helical parameters—pitch, residues per turn, radius, and handedness—and p = rmsd/(N ? 1)1/2 estimating helical regularity, where “rmsd” is the root mean square deviation from the best fit helix and “N” is helix length. A total of 1,496 regular α-helices 6–9 residues long with p ≤ 0.10 Å were identified from 866 protein chains. The statistical analysis provides a strong evidence that the frequency distribution of helices versus n indicates the bimodality of typical α-helix and ω-helix. Sixty-two right handed ω-helices identified (7.2% of proteins) show non-planarity of the peptide groups. There is amino acid preference of Asp and Cys. These observations and analyses insist that the ω-helices occur really in proteins.  相似文献   

3.
Fatty acid hydroperoxide lyase (HPL) is a member of the cytochrome P450 family acting on fatty acid hydroperoxides in many organisms. The active green bell pepper HPL, cloned and expressed in the yeast Yarrowia lipolytica, was purified by immobilized metal-ion affinity chromatography (IMAC) in the presence of 2% of Triton X-100R. The secondary structure prediction by bioinformatics servers of HPL was realized by ANTHEPROT software, using the GOR, DPM and Predator methods. The theoretical results which are average values obtained from three different calculation methods showed 33% α-helix, 18% β-sheet, 7% turn and 42% coil. On the other hand, the secondary structure approach of the purified active HPL (specific activity of 2.94 U/mg protein) was realized by differential scanning calorimetry (DSC) and circular dichroism (CD) spectroscopy, and showed 13% α-helix, 29% β-sheet, 5% turn and 53% random coil.  相似文献   

4.
Structure of the three-chain unit of the bovine epidermal keratin filament   总被引:23,自引:0,他引:23  
The characteristic α-type X-ray diffraction pattern displayed by bovine epidermal keratin filaments can be ascribed to the presence of segments of triple-chain coiled coil α-helix in the repeating three-chain unit of the filaments.Limited proteolysis of filaments polymerized in vitro or a citrate-soluble protein derived from them with crystalline trypsin releases two types of α-helix-enriched particles which provide information on the structure of the three-chain unit. The smaller, particle 2, of molecular weight 42,500, α-helix content of 92% and dimensions of 180 Å × 20 Å, consists of three chains aligned side-by-side that presumably form a coiled coil. The high yields of particle 2 allow the conclusion that all of the α-helix of the epidermal keratin filament is present in the form of these discrete three-chain α-helical segments. The larger, particle 1, recovered during the earlier stages of digestion has a molecular weight of 100,000 to 110,000, α-helix content of 75%, average dimensions of 400 Å × 20 Å and also consists of three chains aligned side-by-side. It contains two α-helical segments corresponding to particle 2 which are located at the amino -terminal and carboxyl-terminal ends and are separated by a region of non-helix. Particle 1 contains all of the α-helix and therefore is the major portion of the three-chain unit of the keratin filament. The products resulting from reaction of intact filament subunits with N-bromosuccinimide suggest that particle 1 is formed during digestion by removal of regions of non-helix from each end of this unit.The structure of the three-chain unit of the bovine epidermal keratin filament may thus be viewed as three polypeptide subunits aligned side-by-side with two discrete coiled coil α-helical segments interspersed with regions of non-helix.  相似文献   

5.
A statistical mechanical model of protein conformation with medium-range interactions between theith and (i+k)th residues (k<-4) is presented. Two two-state models, an α-helix-coil and an extended-structure-coil model, are formulated using the same form of the partition function, but the two models are applied independently to predict the locations of α-helical, extended, and coil segments; in the relatively few cases (<2%) where the predictions from the two models are in conflict, the prediction is scored as an incorrect one. Two independent sets of statistical weights (one set for each model) are derived to describe the interactions between the 20 amino acid residues for each range of interactionk; they are evaluated by minimizing an objective function so that the probability profiles for the α-helix or extended structure, respectively, in proteins computed from these statistical weights correlate optimally with the experimentally observed native conformations of these proteins. Examination of the resulting statistical weights shows that those for the interactions between hydrophobic residues and between a hydrophobic and a hydrophilic residue have reasonable magnitudes compared to what would be expected from the spatial arrangements of the side chains in the α-helix and the extended structure, and that those for the α-helix-coil model correlate well with experimentally determined values of the Zimm-Bragg parameterss and σ of the helix-coil transition theory. From the point of view of a method to predict the conformational states (i.e., α-helix, extended structure, and coil) of each residue, the statistical weights (as inall empirical prediction schemes) depend very much on the proteins used for the data base, since the presently available set of proteins of known structure is still too small for very high predictability; as a result, the correctness of the prediction is not very good for proteins not included in the data base. However, the correctness of the prediction, at least for the 37 proteins utilized as the data base in this study, is 91% and 87% for the α-helix-coil and the extended-structure-coil models, respectively; further, 79% of all the residues are predicted correctly when both the α-helix-coil and extended-structure-coil models are applied independently.  相似文献   

6.
A survey has been made, by using computer methods, of the types of helices which polypeptide chains can form, taking into account steric requirements and intramolecular hydrogen-bonding interactions. The influence on these two requirements, of small variations in the bond angles of the peptide residues, or of small changes in the overall dimensions of the helix (pitch and residues per turn), have been assessed for the special case of the α-helix. Criteria for the formation of acceptable hydrogen bonds have also been applied to helices of other types, viz., the 3, γ?, ω?, and π-helices. It was shown that the N? H … O and H … O? C angles in hydrogen bonds are sensitive to changes in either the NCαC′ bond angle or in the rotational angles about the N? Cα and Cα? C′ bonds. However, the variants of the α-helix observed experimentally in myoglobin can all be constructed without distortion of the hydrogen bonds. For α-helices, the steric and hydrogen bonding requirements are more easily fulfilled with an NCαC′ bond angle of 111°, rather than 109.5°. The decreased stability observed for the left-handed α-helix relative to the right-handed one for L -amino acids is due essentially only to interactions of the Cβ atom of the side chains with atoms in adjacent peptide units in the backbone, and interactions with atoms in adjacent turns of the helical backbone are not significantly different in the two helices. Restrictions in the freedom of rotation of bulky side chains may have significant kinetic effects during the formation of the α-helix from the “random coil” state.  相似文献   

7.
Recently we have developed a new approach to study protein–protein interactions using Fourier transform infrared spectroscopy in combination with titration experiments and principal component analysis (FTIR-TPCA). In the present paper we review the FTIR-TPCA results obtained for the interaction between cytochrome P450 and the redox partner protein in two P450 systems, the Pseudomonas putida P450cam (CYP101) with putidaredoxin (P450cam–Pdx), and the Bacillus megaterium P450BM-3 (CYP102) heme domain with the FMN domain (P450BMP–FMND). Both P450 systems reveal similarities in the structural changes that occur upon redox partner complex formation. These involve an increase in β-sheets and α-helix content, a decrease in the population of random coil/310-helix structure, a redistribution of turn structures within the interacting proteins and changes in the protonation states or hydrogen-bonding of amino acid carboxylic side chains. We discuss in detail the P450cam–Pdx interaction in comparison with literature data and conclusions drawn from experiments obtained by other spectroscopic techniques. The results are also interpreted in the context of a 3D structural model of the Pdx–P450cam complex.  相似文献   

8.
A series of model peptides containing α-trifluoromethyl-substituted amino acids in five different positions relative to the predominant cleavage site of the serine protease α-chymotrypsin was synthesized by solution methods to investigate the influence of α-Tfm substitution on the proteolytic stability of peptides. Proteolysis studies demonstrated absolute stability of peptides substituted in the P1 position and still considerable proteolytic stability for peptides substituted at the P2 and P′2 positions compared with the corresponding unsubstituted model peptide. Comparison with peptides containing the fluorine-free disubstituted amino acid α-aminoisobutyric acid allowed to separate electronic from steric effects. Furthermore, the absolute configuration of the α-Tfm-substituted amino acid was found to exert considerable effects on the proteolytic stability, especially in P′1 substituted peptides. Investigations of this phenomenon using empirical force field calculations revealed that in the (S,R,S)-diasteromer the steric constraints exhibited by the α-Tfm group can be outweighed by an advantageous interaction of the fluorine atoms with the serine side chain of the enzyme. In contrast, a favourable interaction between substrate and enzyme is impossible for the (S,S,S)-diastereomer. © 1997 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

9.
Circular dichroism spectra of the native fatty acid synthetase complex from the insect Ceratitis capitata and of the lipidated and cholate- and SDS-treated enzyme have been obtained. Native enzyme has a calculated structure of 43% α-helix, 23% β structure and 31% random coil. Lipidation and cholate-treatment did not modify the structure of the enzyme complex whereas the SDS-treatment changed the native conformation into a structure based on 42.8% α-helix, 8.4% β structure and 48.8% random coil. These data are interpreted in terms of both the enzyme activity and the quaternary structure of the complex.  相似文献   

10.
A statistical approach has been applied to analyse primary structure patterns at inner positions of α-helices in proteins. A systematic survey was carried out in a recent sample of non-redundant proteins selected from the Protein Data Bank, which were used to analyse α-helix structures for amino acid pairing patterns. Only residues more than three positions apart from both termini of the α-helix were considered as inner. Amino acid pairings i, i+k (k=1, 2, 3, 4, 5), were analysed and the corresponding 20×20 matrices of relative global propensities were constructed. An analysis of (i, i+4, i+8) and (i, i+3, i+4) triplet patterns was also performed. These analysis yielded information on a series of amino acid patterns (pairings and triplets) showing either high or low preference for α-helical motifs and suggested a novel approach to protein alphabet reduction. In addition, it has been shown that the individual amino acid propensities are not enough to define the statistical distribution of these patterns. Global pair propensities also depend on the type of pattern, its composition and orientation in the protein sequence. The data presented should prove useful to obtain and refine useful predictive rules which can further the development and fine-tuning of protein structure prediction algorithms and tools.  相似文献   

11.
AaCtx is the first chlorotoxin-like peptide isolated from Androctonus australis scorpion venom. Its amino acid sequence shares 70% similarity with chlorotoxin from Leiurus quinquestriatus scorpion venom, from which it differs by twelve amino acids. Due to its very low concentration in venom (0.05%), AaCtx was chemically synthesized. Both native and synthetic AaCtx were active on invasion and migration of human glioma cells. However, their activity was found to be lower than that of chlorotoxin. The molecular model of AaCtx shows that most of amino acids differing between AaCtx and chlorotoxin are localized on the N-terminal loop and the α-helix. Based on known compounds that block chloride channels, we suggest that the absence of negative charged amino acids on AaCtx structure may be responsible for its weak activity on glioma cells migration and invasion. This finding serves as a starting point for structure-function relationship studies leading to design high specific anti-glioma drugs.  相似文献   

12.
An isopullulanase (IPU) from Aspergillus niger ATCC9642 hydrolyzes α-1,4-glucosidic linkages of pullulan to produce isopanose. Although IPU does not hydrolyze dextran, it is classified into glycoside hydrolase family 49 (GH49), major members of which are dextran-hydrolyzing enzymes. IPU is highly glycosylated, making it difficult to obtain its crystal. We used endoglycosidase Hf to cleave the N-linked oligosaccharides of IPU, and we here determined the unliganded and isopanose-complexed forms of IPU, both solved at 1.7-Å resolution. IPU is composed of domains N and C joined by a short linker, with electron density maps for 11 or 12 N-acetylglucosamine residues per molecule. Domain N consists of 13 β-strands and forms a β-sandwich. Domain C, where the active site is located, forms a right-handed β-helix, and the lengths of the pitches of each coil of the β-helix are similar to those of GH49 dextranase and GH28 polygalacturonase. The entire structure of IPU resembles that of a GH49 enzyme, Penicillium minioluteum dextranase (Dex49A), despite a difference in substrate specificity. Compared with the active sites of IPU and Dex49A, the amino acid residues participating in subsites + 2 and + 3 are not conserved, and the glucose residues of isopanose bound to IPU completely differ in orientation from the corresponding glucose residues of isomaltose bound to Dex49A. The shape of the catalytic cleft characterized by the seventh coil of the β-helix and a loop from domain N appears to be critical in determining the specificity of IPU for pullulan.  相似文献   

13.
We previously proposed an original two-state cholesterol binding mechanism by StAR, in which the C-terminal α-helix of StAR gates the access of cholesterol to its binding site cavity. This cavity, which can accommodate one cholesterol molecule, was proposed to promote the reversible unfolding of the C-terminal α-helix and allow for the entry and dissociation of cholesterol. In our molecular model of the cholesterol–StAR complex, the hydrophobic moiety of cholesterol interacts with hydrophobic amino acid side-chains located in the C-terminal α-helix and at the bottom of the cavity. In this study, we present a structural in silico analysis of StAR. Molecular dynamics simulations showed that point mutations of Phe267, Leu271 or Leu275 at the α-helix 4 increased the gyration radius (more flexibility) of the protein's structure, whereas the salt bridge double mutant E169M/R188M showed a decrease in flexibility (more compactness). Also, in the latter case, an interaction between Met169 and Phe267 disrupted the hydrophobic cavity, rendering it impervious to ligand binding. These obtained results are in agreement with previous in vitro experiments, and provide further validation of the two-state binding mode of action.  相似文献   

14.
A thermodynamic model describing formation of α-helices by peptides and proteins in the absence of specific tertiary interactions has been developed. The model combines free energy terms defining α-helix stability in aqueous solution and terms describing immersion of every helix or fragment of coil into a micelle or a nonpolar droplet created by the rest of protein to calculate averaged or lowest energy partitioning of the peptide chain into helical and coil fragments. The α-helix energy in water was calculated with parameters derived from peptide substitution and protein engineering data and using estimates of nonpolar contact areas between side chains. The energy of nonspecific hydrophobic interactions was estimated considering each α-helix or fragment of coil as freely floating in the spherical micelle or droplet, and using water/cyclohexane (for micelles) or adjustable (for proteins) side-chain transfer energies. The model was verified for 96 and 36 peptides studied by 1H-nmr spectroscopy in aqueous solution and in the presence of micelles, respectively ([set I] and [set 2]) and for 30 mostly α-helical globular proteins ([set 3]). For peptides, the experimental helix locations were identified from the published medium-range nuclear Overhauser effects detected by 1H-nmr spectroscopy. For sets 1, 2, and 3, respectively, 93, 100, and 97% of helices were identified with average errors in calculation of helix boundaries of 1.3, 2.0, and 4.1 residues per helix and an average percentage of correctly calculated helix—coil states of 93, 89, and 81%, respectively. Analysis of adjustable parameters of the model (the entropy and enthalpy of the helix—coil transition, the transfer energy of the helix backbone, and parameters of the bound coil), determined by minimization of the average helix boundary deviation for each set of peptides or proteins, demonstrates that, unlike micelles, the interior of the effective protein droplet has solubility characteristics different from that for cyclohexane, does not bind fragments of coil, and lacks interfacial area. © 1997 John Wiley & Sons, Inc. Biopoly 42: 239–269, 1997  相似文献   

15.
《FEBS letters》1985,181(1):173-178
The BChlc polypeptide was isolated from chlorosomes of the green bacterium Chloroflexus aurantiacus on Sephadex LH-60. The complete amino acid sequence of this 5.6 kDa polypeptide (51 amino acid residues) was determined. Most probably the 5.6 kDa polypeptide forms an α-helix between Trp 5 and Ile 42 with an asymmetrical (bipolar) distribution of polar amino acid residues along the helix axis: (i) At one side of the α-helix 5 Gln and 2 Asn residues are the possible binding sites for 7 BChlc molecules, (ii) On the other side Ser, Thr, His residues seem to be polypeptide-polypeptide interaction sites within the BChlc-protein complexes. It appears that the BChl-protein complex (chlorosome subunit, 5.2 × 6 nm) composed of 12 5.6 kDa polypeptides corresponds to the 'globular units' found by electron microscopy within the chlorosomes.  相似文献   

16.
17.
The α-helix content of the cytochrome P-450 incorporated into the liposomal membranes of either phosphatidylcholine or microsomal phospholipid insignificantly differed from that of the soluble one. The binding of both type I and type II substrates with cytochrome P-450 incorporated into phosphatidylcholine and microsomal phospholipid membranes did not change the conformation of the polypeptide chain. In contrast to this the type II substrates increased the α-helix content of soluble hemoprotein about 3–5%. Dithionite reduction of the cytochrome P-450 haem increased the degree of α spiralization up to 10% for soluble hemoprotein and up to 5% for the membrane-bound enzyme only. The investigation of the thermal stability of the soluble and liposomal forms of cytochrome P-450 showed that the enzyme incorporated into phospholipid vesicles was highly stable. The heating of the enzyme was followed by a slightly pronounced cooperative transition in contrast to the well-pronounced transition for the soluble cytochrome P-450. Hence, the incorporation of the soluble cytochrome P-450 into phospholipid bilayer does not result in significant change of α-helix content, but the increasing of rigidity and thermal stability of the membrane-bound hemoprotein molecule is observed.  相似文献   

18.
19.
Poly(N-δ-trimethyl-l-ornithine), (Me3Orn)n, is usually not able to attain the α-helical conformation in aqueous solution independent of its pH value; however, it becomes α-helical at low concentrations of sodium perchlorate over a wide pH range according to the circular dichorism (c.d.) spectra. Cl?, SO42? and H2PO4? do not induce α-helix formation. One can conclude that a distinct topology of the anions bound by the side chains is responsible for the α-helix-inducing effect of some water-structure-breaking anions such as perchlorate. This means that the anions are inserted between the ?N+ of the side groups shielding the positive charges repelling one another. The insertion of the anions requires that the water molecules surrounding the ions can be stripped off, which is easily possible if they are water-structure-breaking ones. At higher perchlorate concentrations, the c.d. spectrum changes. It is characterized by a negative shoulder near 208 nm and a pronounced minimum at ≈ 226 nm. With increasing temperature, the c.d. spectrum of the α-helix occurs. Finally the α-helix undergoes a conformational change to the random coil. The apparent transition enthalpy ΔHvH is remarkably lower than that of the homologue (Me3Lys)n, obviously due to a lower cooperativity of the transition. In contrast to poly(l-ornithine), (Orn)n, the c.d. spectrum of (Me3Orn)n remains almost unchanged after adding anionic surfactants such as sodium octyl sulphate (SOS) or sodium dodecyl sulphate (SDS). In organic solvents like methanol or isopropanol, in contrast to (Orn)a and (Lys)n, no α-helix formation occurs. However, in mixtures of these alcohols or dioxane with water, α-helix formation is induced by perchlorate, as in pure water. The thermal stability of the α-helix in these systems is increased.  相似文献   

20.
Plasmodium falciparum malaria continues being one of the parasitic diseases causing the highest worldwide mortality due to the parasite’s multiple evasion mechanisms, such as immunological silence. Membrane and organelle proteins are used during invasion for interactions mediated by high binding ability peptides (HABPs); these have amino acids which establish hydrogen bonds between them in some of their critical binding residues. Immunisation assays in the Aotus model using HABPs whose critical residues had been modified have revealed a conformational change thereby enabling a protection-inducing response. This has improved fitting within HLA-DRβ11 molecules where amino acid electron-donor atoms present in β-turn, random or distorted α-helix structures preferentially bound to HLA-DR53 molecules, whilst HABPs having amino acid electron-acceptor atoms present in regular α-helix structure bound to HLA-DR52. This data has great implications for vaccine development.  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