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 46 毫秒
1.
A theoretical study of effects of excluded volume intermolecular interactions on the sharpness of helix–coil transitions in solutions of polyamino acids or simple proteins indicates that the transition width may vary appreciably as a function of polymer concentration. The analysis is based on a second virial approximation for the excess free energy of mixing of a solution of polymers of varying degrees of helicity. The virial coefficients involved are roughly estimated on the basis of gross polymer geometry. For large N (degree of polymerization) the transition is found, typically to sharpen with increasing concentration, becoming second order and then first order at sufficiently high concentrations. The critical polymer concentration is found to be roughly of the order N?1.2 ??0?1 for an “all or none” model and of order σ1/2 N?0.2 ??0?1 for a model with continuously variable degree of helicity (??0 is the volume of a single helical molecule and σ1/2 the normalized statistical weight of a helix–coil interface). In the second case for N ~ 103 and σ ~ 10?2–10?4, the predicted critical concentration is in the range 10?1–10?3 g/cm.3 Comparison is made with experiments on solutions of poly(γ-benzyl-L glutamate).  相似文献   

2.
Structural characterization of intrinsically disordered proteins (IDPs) is mandatory for deciphering their potential unique physical and biological properties. A large number of circular dichroism (CD) studies have demonstrated that a structural change takes place in IDPs with increasing temperature, which most likely reflects formation of transient α-helices or loss of polyproline II (PPII) content. Using three IDPs, ACTR, NHE1, and Spd1, we show that the temperature-induced structural change is common among IDPs and is accompanied by a contraction of the conformational ensemble. This phenomenon was explored at residue resolution by multidimensional NMR spectroscopy. Intrinsic chemical shift referencing allowed us to identify regions of transiently formed helices and their temperature-dependent changes in helicity. All helical regions were found to lose rather than gain helical structures with increasing temperature, and accordingly these were not responsible for the change in the CD spectra. In contrast, the nonhelical regions exhibited a general temperature-dependent structural change that was independent of long-range interactions. The temperature-dependent CD spectroscopic signature of IDPs that has been amply documented can be rationalized to represent redistribution of the statistical coil involving a general loss of PPII conformations.  相似文献   

3.
Exendin-4 is a natural, 39-residue peptide first isolated from the salivary secretions of a Gila Monster (Heloderma suspectum) that has some pharmacological properties similar to glucagon-like-peptide-1 (GLP-1). This paper reports differences in the structural preferences of these two peptides. For GLP-1 in aqueous buffer (pH 3.5 or 5.9), the concentration dependence of circular dichroism spectra suggests that substantial helicity results only as a consequence of helix bundle formation. In contrast, exendin-4 is significantly helical in aqueous buffer even at the lowest concentration examined (2.3 microM). The pH dependence of the helical signal for exendin-4 indicates that helicity is enhanced by a more favorable sequence alignment of oppositely charged sidechains. Both peptides become more helical upon addition of either lipid micelles or fluoroalcohols. The stabilities of the helices were assessed from the thermal gradient of ellipticity (partial differential theta(221)/partial differential T values); on this basis, the exendin helix does not melt appreciably until temperatures significantly above ambient. The extent of helix formation for exendin-4 in aqueous buffer (and the thermal stability of the resulting helix) suggests the presence of a stable helix-capping interaction which was localized to the C-terminal segment by NMR studies of NH exchange protection. Solvent effects on the thermal stability of the helix indicate that the C-terminal capping interaction is hydrophobic in nature. The absence of this C-capping interaction and the presence of a flexible, helix-destabilizing glycine at residue 16 in GLP-1 are the likely causes of the greater fragility of the monomeric helical state of GLP-1. The intramolecular hydrophobic clustering in exendin-4 also appears to decrease the extent of helical aggregate formation.  相似文献   

4.
Earlier studies have shown that the helical content of α‐helical peptide decreases upon its interaction with carbon nanotube (CNT). Further, the length of the α‐helix varies from few residues in the small globular protein to several number of residues in structural and membrane proteins. In structural and membrane proteins, helices are widely present as the supercoil i.e., helical bundles. Thus, in this study, the length‐dependent interaction pattern of α‐helical peptides with CNT and the stability of isolated α‐helical fragment versus supercoiled helical bundle upon interaction with CNT have been investigated using classical molecular dynamics (MD) simulation. Results reveal that the disruption in the helical motif on interaction with CNT is directly proportional to the length of the helix. Also it is found that the shorter helix does not undergo noticeable changes in the helicity upon adsorption with CNT. On the other hand, helicity of longer peptides is considerably affected by its interaction with CNT. In contrast to the known fact that the stability of the helix increases with its length, the disruption in the helical peptide increases with its length upon its interaction with CNT. Comparison of results shows that structural changes in the isolated helical fragment are higher than that in supercoiled helix. In fact, helical chain in supercoiled bundle does not undergo significant changes in the helicity upon interaction with CNT. Both the length of the helical peptide and the inherent stability of the helical unit in the supercoiled helix influence the interaction pattern with the CNT. © 2012 Wiley Periodicals, Inc. Biopolymers 99: 357–369, 2013.  相似文献   

5.
Properties of a simple model of polypeptide chains were studied by the means of the Monte Carlo method. The chains were built on the (310) hybrid lattice. The residues interacted with long-range potential. There were two kinds of residues: hydrophobic and hydrophilic forming a typical helical pattern -HHPPHPP-. Short range potential was used to prefer helical conformations of the chain. It was found that at low temperatures the model chain formes dense and partially ordered structures (non-unique). The presence of the local potential led to an increase of helicity. The effect of the interplay between the two potentials was studied. After the collapse of the chain further annealing caused rearrangement of helical structures. Dynamic properties of the chain at low temperature depended strongly on the local chain ordering.  相似文献   

6.
7.
The temperature dependence of helical propensities for the peptides Ac-ZGG-(KAAAA)(3)X-NH(2) (Z = Y or G, X = A, K, and D-Arg) were studied both experimentally and by MD simulations. Good agreement is observed in both the absolute helical propensities as well as relative helical content along the sequence; the global minimum on the calculated free energy landscape corresponds to a single alpha-helical conformation running from K4 to A18 with some terminal fraying, particularly at the C-terminus. Energy component analysis shows that the single helix state has favorable intramolecular electrostatic energy due to hydrogen bonds, and that less-favorable two-helix globular states have favorable solvation energy. The central lysine residues do not appear to increase helicity; however, both experimental and simulation studies show increasing helicity in the series X = Ala --> Lys --> D-Arg. This C-capping preference was also experimentally confirmed in Ac-(KAAAA)(3)X-GY-NH(2) and (KAAAA)(3)X-GY-NH(2) sequences. The roles of the C-capping groups, and of lysines throughout the sequence, in the MD-derived ensembles are analyzed in detail.  相似文献   

8.
An efficient Monte Carlo (MC) algorithm using concerted backbone rotations is combined with a recently developed implicit membrane model to simulate the folding of the hydrophobic transmembrane domain M2TM of the M2 protein from influenza A virus and Sarcolipin at atomic resolution. The implicit membrane environment is based on generalized Born theory and has been calibrated against experimental data. The MC sampling has previously been used to fold several small polypeptides and been shown to be equivalent to molecular dynamics (MD). In combination with a replica exchange algorithm, M2TM is found to form continuous membrane spanning helical conformations for low temperature replicas. Sarcolipin is only partially helical, in agreement with the experimental NMR structures in lipid bilayers and detergent micelles. Higher temperature replicas exhibit a rapidly decreasing helicity, in agreement with expected thermodynamic behavior. To exclude the possibility of an erroneous helical bias in the simulations, the model is tested by sampling a synthetic Alanine-rich polypeptide of known helicity. The results demonstrate there is no overstabilization of helical conformations, indicating that the implicit model captures the essential components of the native membrane environment for M2TM and Sarcolipin.  相似文献   

9.
S Honda  S Ohashi  H Morii  H Uedaira 《Biopolymers》1991,31(7):869-876
The conformations of synthetic human growth hormone-releasing factor fragment (1-29) in the presence and the absence of 1,2-dimyristoyl-sn-glycero-3-phosphorylglycerol liposome as well as in aqueous 2,2,2-trifluoroethanol solution were investigated by CD spectroscopy. The secondary structure of the peptide in each solution was analyzed by two methods. Both results show that the peptide has an unordered structure in the aqueous solution, whereas it folds into helical structure in the aqueous alcohol and in the phospholipid solution. In addition, although the peptide exists as almost complete helix in the 50 vol% aqueous alcohol (80-90% helicity), it does not reach full helicity even in the solution containing excess amount of phospholipid liposome (maximum 65-70% helicity). The conformational difference is explained by the characteristic amphipathy of the peptide, i.e., the necessity to twist the separated amphipathic helical parts in the interaction with the phospholipid membrane probably makes the helicity of the peptide decrease.  相似文献   

10.
Replica exchange molecular dynamics simulations are used to generate three ensembles of an S-peptide analog (AETAAAKFLREHMDS). Percent helicity of the peptide ensembles calculated using STRIDE is compared to percent helicity calculated from (13)C(alpha) chemical shift deviations (CSD) from random coil in order to test the assumption that CSD can be correlated to percent helicity. The two estimates of helicity, one based on structure and the other on CSD, are in close to quantitative agreement, except at the edges of helical stretches where disagreements of as much as 50% can be found. These disagreements can occur by CSDs both as an under- and an overestimate of peptide helicity. We show that underestimation arises due to ensemble averaging of positive CSDs from conformers with torsion angles in the helical region of Ramachandran space with negative CSDs corresponding to conformers of the peptide in the extended region. In contrast, overestimation comes about due to the fact that a large number of conformations with torsion angles in the helical region are not counted as helical by STRIDE due to a lack of correlated helical torsion angles in neighboring residues.  相似文献   

11.
The reformation of secondary structure for unfolded, disulfide reduced hen egg white lysozyme (HEWL) upon interaction with surfactants was studied using CD, fluorescence and IR (infrared) techniques. Equilibrium CD studies showed that reduced HEWL when mixed with negatively charged surfactants, such as SDS (sodium dodecyl sulfate), gradually regains average helical structure to a level equivalent to that obtained for the oxidized form also in SDS, but both forms lose tertiary structure in such environments. This non-native structure recovery process begins with monomer surfactant interaction but at higher concentrations is in part dependent on micelle formation, with the helical fraction reaching its maximum value with each surfactant only above the CMC. Fluorescence changes were more complex, evidencing an intermediate state at lower surfactant concentration. With positively charged surfactants the degree of helicity recovered was less, and the intermediate state in fluorescence was not seen. Stopped flow dynamics studies showed the CD kinetics fit to two exponentials as did the fluorescence. The faster steps in CD and fluorescence detected kinetics appear to be correlated which suggests formation of an intermediate on rapid interaction of the micelle and protein. The second step then reflected attainment of a stable surfactant solvated state which attains maximum helicity and moves the Trps to a more hydrophobic environment, which may occur in independent steps, as the slower kinetics are not well correlated.  相似文献   

12.
A 17-residue O-linked glycopeptide model incorporating a central alpha-mannosyl serine residue, and its unglycosylated analog both demonstrate substantial helicity in water. The peptide sequence was derived from previous studies in which differences in overall helicity as a function of single amino acid substitutions were measured by circular dichroism (CD). The helical content was predicted by molecular modeling, and confirmed by CD and NMR. Moreover, the glycopeptide retained its helicity in the presence of SDS micelles, whereas the native peptide lost secondary structure in the presence of micelles. The inference is that the peptide sequence is a more important helix determinant than glycosylation per se.  相似文献   

13.
A Yaron  N Tal  A Berger 《Biopolymers》1972,11(12):2461-2481
The sequence-ordered copolymer poly-(Lys-Ala-Ala) was synthesized by polycondensation of the N-hydroxysuccinimide ester of ε,Z-Lys-Ala-Ala and deprotection of the polymerization product. A fraction of molecular weight 13,000 obtained by ion-exchange chromatography was investigated. The polymer is freely soluble in water at all pH values, and is completely digested by trypsin and elastase. From CD and ORD data it was concluded that in water at 1°C the ionized form (at pH 6.5) of the polymer is helical. On heating, helix-coil transition curves were obtained with a midpoint, Tm, depending on salt concentration. In salt-free water Tm = 12.3°C and in 0.2M NaCl Tm = 28.5°C. Adding MeOH, causes an increase in the helical content of the polymer (half helicity at 20% MeOH, without salt, at 29°C). Guanidine·HCl was shown to decrease the helicity. At 1°C half helicity. The nonionized polymer helix is more stable (Tm~90°C). At the high pH, at 60°C, when concentration of the polymer is higher than 1.9 × 10-2M, a precipitate is formed which redissolves on cooling with the original helicity. This does not occur in the presence of 50% MeOH. By comparison with polylysine it was concluded that replacing two-thirds of the lysine residues in polylysine by alanine leads to a polymer forming a more stable α-helix, when fully ionized. This is essentially due to the diminished coulombic repulsion. Uncharged lysine residues are comparable to alanine residues in their helix-forming tendency since the sequential polymer as well as one-third ionized polylysine are helical to approximately the same extent at room temperature.  相似文献   

14.
The effect of salts on the experimental circular dichroism spectra of polypeptides is presented using poly-L-lysine as the main model. Salt effects are analyzed into: (a) shielding at low (less than 0.5 M) concentrations of all salts; (b) binding to positively charged and some neutrally charged side-chains by certain anions (e.g., CCl3COO-, CF3C00-, ClO4-), with induction of helicity; (c) binding of these same anions, at high concentration, to the backbone leading toward random structure; (d) binding of high concentration of denaturing cations (La+3, Ca++, Li+) to the backbone, with La+3 and Ca++ leading to collapsed random structure (R) while Li+ tends to leave the polypeptide somewhat extended; (e) indirect interaction of salting-out salts (NaH2PO4, (NH4)2SO4, NH4F), at high concentration, leading toward complete alpha helicity, probably by competition with the polypeptide and the anion for available water. Effects of changing the temperature from 5 degrees to 50 degrees on the circular dishroism spectra of different polypeptide-salt solutions throughout the region from extended (LES) to alpha helical conformation are analyzed in terms of introduction of randomness (R) at high temperature. Applications to effects of salt on protein structures are considered.  相似文献   

15.
The optical rotatory dispersion (ORD), and to a lesser extent the circular dichroism (CD) of the proteins tropomyosin and myosin have been extensively studied. The effect of aging and of certain reagents on these optical properties, relating to helical contents, were observed. As expected, spectra typical of right-handed, α-helices were found, with helical contents of 60–65% for myosin and 97–100% for tropomyosin. Studies were made to find which aqueous salt solutions would permit rotation observations in the far ultraviolet and also dissolve these proteins. The presence or absence of sulfhydryl groups (SH) were found to have no effect on helicity, and it is suggested that both proteins have an identical helical portion of 40%.  相似文献   

16.
Gad-1 and Gad-2 are antimicrobial peptide (AMP) sequences encoded by paralogous genes. They are rich in histidine, which suggests that their activity might be pH-dependent. We examined their structure–function relationships with a view to learning how to improve AMP therapeutic ratios. Activity assays with Gram-negative bacteria and cancer cell lines demonstrate that Gad-2 is substantially more active at slightly acidic pH than it is at neutral pH. By contrast, the activity of Gad-1 at lower pH is similar to its activity at pH 7. Circular dichroism spectra indicate that the greater functional plasticity of Gad-2 correlates with a greater structural plasticity; Gad-2's percent helicity varies dramatically with altered pH and lipid environment. Interestingly, Gad-2's highest levels of helicity do not correspond to the conditions where it is most active. High resolution solution NMR structures were determined in SDS micelles at pH 5, conditions that induce an intermediate level of helicity in the peptides. Gad-1 is more helical than Gad-2, with both peptides exhibiting the greatest helical tendencies in their central region and lowest helicity in their N-termini. The high resolution structures suggest that maximum activity relies on the appropriate balance between an N-terminal region with mixed hydrophobic/hydrophilic structure features and an amphipathic central and C-terminal region. Taken together with previous studies, our results suggest that to improve the therapeutic ratio of AMPs, consideration should be given to including sequential histidine-pairs, keeping the overall charge of the peptide modest, and retaining a degree of structural plasticity and imperfect amphipathicity.  相似文献   

17.
Poly-N-substituted glycines or "peptoids" are protease-stable peptide mimics. Although the peptoid backbone is achiral and lacks hydrogen-bond donors, substitution with alpha-chiral side chains can drive the formation of stable helices that give rise to intense CD spectra. To systematically study the solution properties and stability of water-soluble peptoid helices with alpha-chiral side chains, we have synthesized and characterized an amphipathic, 36-residue N-substituted glycine oligomer. CD was used to investigate effects of concentration and solvent environment on this helical peptoid. We saw no significant dependence of helical structure on concentration. Intense, "alpha-helix-like" CD spectra were observed for the 36-mer in aqueous, 2,2,2-trifluorethanol (TFE), and methanol solution, proving a relative insensitivity of peptoid helical structure to solvent environment. While CD spectra taken in these different solvents were fundamentally similar in shape, we did observe some interesting differences in the intensities of particular CD bands in the various solvents. For example, the addition of TFE to an aqueous solvent increases the degree of peptoid helicity, as is observed for polypeptide alpha-helices. Moreover, the helical structure of peptoids appears to be virtually unaffected by heat, even in an aqueous buffer containing 8 M urea. The extraordinary resistance of these peptoid helices to denaturation is consistent with a dominant role of steric forces in their structural stabilization. The structured polypeptoids studied here may have potential as robust mimics of helical polypeptides of therapeutic interest.  相似文献   

18.
Extent of Double Strandedness in Avian Myeloblastosis Virus RNA   总被引:3,自引:0,他引:3       下载免费PDF全文
The extent of double strandedness of avian myeloblastosis virus 70S RNA has been determined from fluorescence measurements of the intercalation of ethidium bromide. We have shown that 50% of the nucleotides of 70S RNA in solution are in a stable helical configuration. This value does not include small helical regions that are too unstable to permit intercalation of the dye. The avian myeloblastosis virus RNA as it exists within the virion has the same degree of helicity as the free 70S RNA. Heating the free 70S RNA to 55 or 70 C, followed by cooling, does not measurably change the degree of helicity; the subunits therefore have as much helicity as the parent molecule.  相似文献   

19.
Potentiometric titrations and some complementary optical rotation data are presented for solutions of poly(L - glutamic acid) (PGA) in several H2O–ethanol mixtures. The data allow the determination of the intrinsic pK (pK0), slope of the apparent. pK (pKapp), versus degree of ionization curves and of the enthalpy of ionization as a function of ethanol concentration. The variation of the degree of ionization at which the helix–coil transformation occurs with ethanol and temperature is also determined. Finally free energy, enthalpy, and intropy changes associated with the helix–coil transformation for the uncharged conformers are determined from the titration curves. The effect of the ethanol is to increase the stability of the helical conformation of PGA for both the charged and the uncharged forms of the polymer. The stabilization of the uncharged helix is essentially an entropic effect.  相似文献   

20.
J F Collawn  Y Paterson 《Biopolymers》1990,29(8-9):1289-1296
The conformations of two 17-residue peptide analogues derived from the C-terminal sequence of pigeon cytochrome c (native sequence = KAERADLIAYLKQATAK) were examined in aqueous and lipid environments by CD spectroscopy. The two analogues, KKLLKKLIAYLKQATAK (K peptide) and EELLEELIAYLKQATAK (E peptide), were made amphipathic with respect to helical segregation by substituting a 6-residue sequence at the N-terminus of the native peptide. Their structures were compared to the native peptide under aqueous conditions of varying pH and temperature, and in the presence of liposomes composed of phosphatidylcholine and phosphatidylserine in the ratio of 9:1. The results indicated that the native peptide remains unstructured under all the conditions examined even though this region of the native molecule is surface exposed and helical. The E peptide, however, was helical under aqueous conditions at 25 degrees C from pH 2-10 with a maximum helicity at pH 4 (54% helix from analysis of CD data). The ellipticity of the E peptide at pH 4 and 8 was concentration dependent, indicating an aggregation phenomenon. In studies in which the CD spectrum was measured at different temperatures, the E peptide became more helical at lower temperatures at pH 4 but not at pH 8. Upon interaction with a lipid membrane in the form of liposomes, there appeared to be a slight destabilization in the structure of the E peptide. The K peptide in an aqueous environment behaved like the native peptide in that it was structureless at all pHs and temperatures examined. In the presence of liposomes, however, this peptide had a high helical content (75% helix from analysis of CD data). These findings suggest that while stabilization of the helix dipole with negative charges at the N-terminus are important in inducing helical conformation in the E peptide, hydrophobic interactions created during aggregation appear to provide the principal stabilizing force. The results with the K peptide demonstrate that the positive N-terminal sequence of this peptide is able to interact with the negatively charged head groups in the phospholipid membrane in such a fashion as to stabilize a helical structure that is not apparent in an aqueous environment alone.  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