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 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   

2.
D J Tobias  C L Brooks 《Biochemistry》1991,30(24):6059-6070
We used molecular dynamics simulations to study the folding/unfolding of one of turn of an alpha helix in Ac-(Ala)3-NHMe and Ac-(Val)3-NHMe. Using specialized sampling techniques, we computed free energy surfaces as functions of a conformational coordinate that corresponds to alpha helices at small values and to extended conformations at large values. Analysis of the peptide conformations populated during the simulations showed that alpha helices, reverse turns, and extended conformations correspond to minima on the free energy surfaces of both peptides. The free energy difference between alpha helix and extended conformations, determined from the equilibrium constants for helix unfolding, is approximately -1 kcal/mol for Ac-(Ala)3-NHMe and -5 kcal/mol for Ac-(Val)3-NHMe. The mechanism observed in our simulations, which includes reverse turns as important intermediates along the helix folding/unfolding pathway, is consistent with a mechanism proposed previously. Our results predict that both peptides (but especially the Ala peptide) have a much larger equilibrium constant for helix initiation than is predicted by the helix-coil transition theory with the host-guest parameters. We also predict a much greater difference in the equilibrium constants than the theory predicts. Insofar as helix initiation is concerned, our results suggest that the large difference between the helical propensities of Ala and Val cannot be explained by simple concepts such as side-chain rotamer restriction or unfavorable steric interactions. Rather, the origin of the difference appears to be quite complicated because it involves subtle differences in the solvation of the two peptides. The two peptides have similar turn-extended equilibria but very different helix-turn equilibria, and the difference in helical propensities reflects the fact that the helix-turn equilibrium strongly favors the turns in Ac-(Val)3-NHMe, while it favors the helices in Ac-(Ala)3-NHMe. We also computed thermodynamic decompositions of the free energy surfaces, and these revealed that the helix-turn equilibria are vastly different primarily because the changes in peptide-water interactions that accompany helix-to-turn conformational changes are qualitatively different for the two peptides.  相似文献   

3.
Nymeyer H  Woolf TB  Garcia AE 《Proteins》2005,59(4):783-790
We implement the replica exchange molecular dynamics algorithm to study the interactions of a model peptide (WALP-16) with an explicitly represented DPPC membrane bilayer. We observe the spontaneous, unbiased insertion of WALP-16 into the DPPC bilayer and its folding into an alpha-helix with a transbilayer orientation. The free energy surface suggests that the insertion of the peptide into the DPPC bilayer precedes secondary structure formation. Although the peptide has some propensity to form a partially helical structure in the interfacial region of the DPPC/water system, this state is not a productive intermediate but rather an off-pathway trap for WALP-16 insertion. Equilibrium simulations show that the observed insertion/folding pathway mirrors the potential of mean force (PMF). Calculation of the enthalpic and entropic contributions to this PMF show that the surface bound conformation of WALP-16 is significantly lower in energy than other conformations, and that the insertion of WALP-16 into the bilayer without regular secondary structure is enthalpically unfavorable by 5-10 kcal/mol/residue. The observed insertion/folding pathway disagrees with the dominant conceptual model, which is that a surface-bound helix is an obligatory intermediate for the insertion of alpha-helical peptides into lipid bilayers. In our simulations, the observed insertion/folding pathway is favored because of a large (>100 kcal/mol) increase in system entropy that occurs when the unstructured WALP-16 peptide enters the lipid bilayer interior. The insertion/folding pathway that is lowest in free energy depends sensitively on the near cancellation of large enthalpic and entropic terms. This suggests the possibility that intrinsic membrane peptides may have a diversity of insertion/folding behaviors depending on the exact system of peptide and lipid under consideration.  相似文献   

4.
Amino acids in peptides and proteins display distinct preferences for alpha-helical, beta-strand, and other conformational states. Various physicochemical reasons for these preferences have been suggested: conformational entropy, steric factors, hydrophobic effect, and backbone electrostatics; however, the issue remains controversial. It has been proposed recently that the side-chain-dependent solvent screening of the local and non-local backbone electrostatic interactions primarily determines the preferences not only for the alpha-helical but also for all other main-chain conformational states. Side-chains modulate the electrostatic screening of backbone interactions by excluding the solvent from the vicinity of main-chain polar atoms. The deficiency of this electrostatic screening model of amino acid preferences is that the relationships between the main-chain electrostatics and the amino acid preferences have been demonstrated for a limited set of six non-polar amino acid types in proteins only. Here, these relationships are determined for all amino acid types in tripeptides, dekapeptides, and proteins. The solvation free energies of polar backbone atoms are approximated by the electrostatic contributions calculated by the finite difference Poisson-Boltzmann and the Langevin dipoles methods. The results show that the average solvation free energy of main-chain polar atoms depends strongly on backbone conformation, shape of side-chains, and exposure to solvent. The equilibrium between the low-energy beta-strand conformation of an amino acid (anti-parallel alignment of backbone dipole moments) and the high-energy alpha conformation (parallel alignment of backbone dipole moments) is strongly influenced by the solvation of backbone polar atoms. The free energy cost of reaching the alpha conformation is by approximately 1.5 kcal/mol smaller for residues with short side-chains than it is for the large beta-branched amino acid residues. This free energy difference is comparable to those obtained experimentally by mutation studies and is thus large enough to account for the distinct preferences of amino acid residues. The screening coefficients gamma(local)(r) and gamma(non-local)(r) correlate with the solvation effects for 19 amino acid types with the coefficients between 0.698 to 0.851, depending on the type of calculation and on the set of point atomic charges used. The screening coefficients gamma(local)(r) increase with the level of burial of amino acids in proteins, converging to 1.0 for the completely buried amino acid residues. The backbone solvation free energies of amino acid residues involved in strong hydrogen bonding (for example: in the middle of an alpha-helix) are small. The hydrogen bonded backbone is thus more hydrophobic than the peptide groups in random coil. The alpha-helix forming preference of alanine is attributed to the relatively small free energy cost of reaching the high-energy alpha-helix conformation. These results confirm that the side-chain-dependent solvent screening of the backbone electrostatic interactions is the dominant factor in determining amino acid conformational preferences.  相似文献   

5.
We have observed that the rate of folding of the enzymatically hydroxylated form of poly(Gly-Pro-Pro) into the triple-helical conformation is considerably higher than that of the unhydroxylated polypeptide [R. K. Chopra and V. S. Ananthanarayanan (1982) Proc. Natl. Acad. Sci. USA 79 , 7180–7184]. In this study, we examine a plausible kinetic pathway for triple-helix formation by selecting peptide models for the unhydroxylated collagen molecule, and computing their conformational energies before and after proline hydroxylation. Starting with the available data on the preferred conformations of proline- and hydroxyproline-containing peptide sequences, energy minimization was carried out on the following pairs of peptides: Gly-Ala-Pro-Gly-Ala and Gly-Ala-Hyp-Gly-Ala; Gly-Pro-Pro-Gly-Ala and Gly-Pro-Hyp-Gly-Ala; Gly-Ala-Pro-Gly-Ala-Pro and Gly-Ala-Hyp-Gly-Ala-Hyp. It was found that, with each pair of peptides, the energetically most favorable conformation (I) has an extended structure at the Gly-Ala or Gly-Pro segment and a β-bend at the Pro-Gly or Hyp-Gly segment. In the Hyp-containing peptides, this conformation is further stabilized by a (Hypi + 2)OH…OC(Glyi) hydrogen bond. Conformation I is lower in energy by about 6–13 kcal/mol of the peptide than the fully extended conformations that resemble the single collagen polypeptide chain and contain no intramolecular hydrogen bond. In contrast to the proline counterpart, the hydroxyproline-containing peptides are found capable of adopting a partially extended conformation that does not contain the β-bend but retains the (Hyp)OH…OC(Gly) hydrogen bond. The energy of this conformation is intermediate between conformation I and the fully extended conformation. The continuation of the β-bend along the chain is restricted by stereochemical constraints that are more severe in the latter two pairs of peptides than in the first pair. Such a restriction may be considered to trigger the “unbending” of the minimum energy conformation leading to its straightening into the fully extended conformation; the latter, in turn, would lead to triple-helix formation through favorable interchain interactions. We propose that the partially extended conformation in the Hyp-containing peptides could serve as a kinetic intermediate on the way to forming the fully extended conformation. Because of the (Hypi + 2)OH…OC(Glyi) hydrogen bond, this conformation would also serve to lock the trans geometry at the Gly-Ala(Pro) and Ala(Pro)-Hyp peptide bonds, thereby enhancing the rate of their helix formation. A scheme for collagen folding in proposed on the basis of these results.  相似文献   

6.
The proline residues in proteins are known to play an important structural role. Recently, the role of a proline residue in the middle of right-handed alpha-helical segments of peptides has been the focus of attention. This role seems to be particularly important in the case of membrane proteins and in the tight packing of globular proteins. In the present study the right-handed alpha-helical region of the Ala-Pro dipeptide and of polypeptides containing this group have been investigated. Crystal structures of proline-containing alpha-helices from some proteins have been analyzed and energy minimization studies on some model fragments containing Ala-Pro in the right-handed alpha-helical conformation have been carried out using flexible geometry. The present calculations indicate that the right-handed alpha-helical region of conformational space is an energetically favored region that can also accommodate Ala-Pro in longer segments of right-handed alpha-helix. This is achieved due to minor variations in some of the internal parameters. Deviations in the backbone parameters of proline in the right-handed alpha-helix lead to a kink of about 23 degrees in the helix axis. These deviations have been characterized and a set of standard values has been suggested for producing such a kink. These values can be used for model building and as starting points for further minimization studies. Previous energy minimization studies have been done using rigid geometry. This may explain why the minimum for Ala-Pro in the right-handed alpha-helical region has not been recognized thus far.  相似文献   

7.
J H Hurley  D A Mason  B W Matthews 《Biopolymers》1992,32(11):1443-1446
Previously calculated conformational energy maps suggest that the alpha-helical conformation for the residue preceding a proline is disfavored relative to the extended conformation by more than 7 kcal/mol. In known protein structures this conformation is observed, however, to occur for about 9% of all prolines. In addition, introduction or removal of prolines at theoretically unfavorable positions in proteins and peptides can have modest effects on stability and structure. To investigate the discrepancy between calculation and experiment, we have determined how the conformation of the proline affects the calculated energy. We have also explored the effect of bond length and bond angle relaxation on the conformational energy map. The conformational energy of the preceding residue is found to be unaffected by the conformation of the proline, but the effect of allowing covalent bond relaxation is dramatic. If bond lengths and angles, and dihedral angles within the pyrrolidine ring, are allowed to relax, a calculated energy difference between the alpha and beta conformations of 1.1 kcal/mol is obtained, in reasonable agreement with experiment. The detailed shape of the calculated energy surface is also in excellent agreement with the observed conformational distributions in known protein structures.  相似文献   

8.
The average globular protein contains 30% alpha-helix, the most common type of secondary structure. Some amino acids occur more frequently in alpha-helices than others; this tendency is known as helix propensity. Here we derive a helix propensity scale for solvent-exposed residues in the middle positions of alpha-helices. The scale is based on measurements of helix propensity in 11 systems, including both proteins and peptides. Alanine has the highest helix propensity, and, excluding proline, glycine has the lowest, approximately 1 kcal/mol less favorable than alanine. Based on our analysis, the helix propensities of the amino acids are as follows (kcal/mol): Ala = 0, Leu = 0.21, Arg = 0.21, Met = 0.24, Lys = 0.26, Gln = 0.39, Glu = 0.40, Ile = 0.41, Trp = 0.49, Ser = 0.50, Tyr = 0. 53, Phe = 0.54, Val = 0.61, His = 0.61, Asn = 0.65, Thr = 0.66, Cys = 0.68, Asp = 0.69, and Gly = 1.  相似文献   

9.
The hydrogen-deuterium exchange kinetics of 37 backbone amide residues in RNase T1 have been monitored at 25, 40, 45, and 50 degrees C at pD 5.6 and at 40 and 45 degrees C at pD 6.6. The hydrogen exchange rate constants of the hydrogen-bonded residues varied over eight orders of magnitude at 25 degrees C with 13 residues showing exchange rates consistent with exchange occurring as a result of global unfolding. These residues are located in strands 2-4 of the central beta-pleated sheet. The residues located in the alpha-helix and the remaining strands of the beta-sheet exhibited exchange behaviors consistent with exchange occurring due to local structural fluctuations. For several residues at 25 degrees C, the global free energy change calculated from the hydrogen exchange data was over 2 kcal/mol greater than the free energy of unfolding determined from urea denaturation experiments. The number of residues showing this unexpected behavior was found to increase with temperature. This apparent inconsistency can be explained quantitatively if the cis-trans isomerization of the two cis prolines, Pro-39 and Pro-55, is taken into account. The cis-trans isomerization equilibrium calculated from kinetic data indicates the free energy of the unfolded state will be 2.6 kcal/mol higher at 25 degrees C when the two prolines are cis rather than trans (Mayr LM, Odefey CO, Schutkowski M, Schmid FX. 1996. Kinetic analysis of the unfolding and refolding of ribonuclease T1 by a stopped-flow double-mixing technique. Biochemistry 35: 5550-5561). The hydrogen exchange results are consistent with the most slowly exchanging hydrogens exchanging from a globally higher free energy unfolded state in which Pro-55 and Pro-39 are still predominantly in the cis conformation. When the conformational stabilities determined by hydrogen exchange are corrected for the proline isomerization equilibrium, the results are in excellent agreement with those from an analysis of urea denaturation curves.  相似文献   

10.
The assumption that the intrinsic alpha-helical propensities of the amino acids are position independent was critical in several helix/coil transition theories. In the first paper of these series, we reported that this is not the case for Gly and nonpolar aliphatic amino acids (Val, Leu, Met, and Ile). Here we have analyzed the helical intrinsic propensities of noncharged polar residues (Ser, Thr, Asn, and Gln) at different positions of a model polyalanine-based peptide. We found that Thr is more favorable (by approximately 0.3 kcal/mol) at positions N1 and N2 than in the helix center, although for Ser, Asn, and Gln the differences are smaller (+/-0.2 kcal/mol), and in many cases within the experimental error. There is a reasonable agreement (+/-0.2 kcal/mol) between the calculated free energies, using the ECEPP/2 force field equipped with a hydration potential, and the experimental data, except at position N1.  相似文献   

11.
Eglin c, turkey ovomucoid third domain, and bovine pancreatic trypsin inhibitor (Kunitz) are all standard mechanism, canonical protein inhibitors of serine proteinases. Each of the three belongs to a different inhibitor family. Therefore, all three have the same canonical conformation in their combining loops but differ in their scaffoldings. Eglin c (Leu45 at P1) binds to chymotrypsin much better than its Ala45 variant (the difference in standard free energy changes on binding is -5.00 kcal/mol). Similarly, turkey ovomucoid third domain (Leu18 at P1) binds to chymotrypsin much better than its Ala18 variant (the difference in standard free energy changes on binding is -4.70 kcal/mol). As these two differences are within the +/-400 cal/mol bandwidth (expected from the experimental error), one can conclude that the system is additive. On the basis that isoenergetic is isostructural, we expect that within both the P1 Ala pair and the P1 Leu pair, the conformation of the inhibitor's P1 side chain and of the enzyme's specificity pocket will be identical. This is confirmed, within the experimental error, by the available X-ray structures of complexes of bovine chymotrypsin Aalpha with eglin c () and with turkey ovomucoid third domain (). A comparison can also be made between the structures of P1 (Lys+)15 of bovine pancreatic trypsin inhibitor (Kunitz) ( and ) and of the P1 (Lys+)18 variant of turkey ovomucoid third domain (), both interacting with chymotrypsin. In this case, the conformation of the side chains is strikingly different. Bovine pancreatic trypsin inhibitor with (Lys+)15 at P1 binds to chymotrypsin more strongly than its Ala15 variant (the difference in standard free energy changes on binding is -1.90 kcal/mol). In contrast, turkey ovomucoid third domain variant with (Lys+)18 at P1 binds to chymotrypsin less strongly than its Ala18 variant (the difference in standard free energies of association is 0.95 kcal/mol). In this case, P1 Lys+ is neither isostructural nor isoenergetic. Thus, a thermodynamic criterion for whether the conformation of a P1 side chain in the complex matches that of an already determined one is at hand. Such a criterion may be useful in reducing the number of required X-ray crystallographic structure determinations. More importantly, the criterion can be applied to situations where direct determination of the structure is extremely difficult. Here, we apply it to determine the conformation of the Lys+ side chain in the transition state complex of a substrate with chymotrypsin. On the basis of kcat/KM measurements, the difference in free energies of activation for Suc-AAPX-pna when X is Lys+ and X is Ala is 1.29 kcal/mol. This is in good agreement with the corresponding difference for turkey ovomucoid third domain variants but in sharp contrast to the bovine pancreatic trypsin inhibitor (Kunitz) data. Therefore, we expect that in the transition state complex of this substrate with chymotrypsin, the P1 Lys+ side chain is deeply inserted into the enzyme's specificity pocket as it is in the (Lys+)18 turkey ovomucoid third domain complex with chymotrypsin.  相似文献   

12.
R H Yun  A Anderson  J Hermans 《Proteins》1991,10(3):219-228
Free-energy simulations have been used to estimate the change in the conformational stability of short polyalanine alpha-helices when one of the alanines is replaced by a proline residue. For substituting proline in the middle of the helix the change in free energy of folding (delta delta G degrees) was calculated as 14 kJ/mol (3.4 kcal/mol), in excellent agreement with the one available experimental value. The helix containing proline was found to be strongly kinked; the free energy for reducing the angle of the kink from 40 degrees to 15 degrees was calculated, and found to be small. A tendency to alternate hydrogen bonding schemes was observed in the proline-containing helix. These observations for the oligopeptide agree well with the observation of a range of kink angles (18-35 degrees) and variety of hydrogen bonding schemes, in the rare instances where proline occurs in helices in globular proteins. For substituting proline at the N-terminus of the helix the change in free energy of folding (delta delta G degrees) was calculated as -4 kJ/mol in the first helical position (N1) and +6 kJ/mol in the second helical position (N2). The observed frequent occurrence of proline in position N1 in alpha-helices in proteins therefore has its origin in stability differences of secondary structure. The conclusion reached here that proline may be a better helix former in position N1 than (even) alanine, and thus be a helix initiator may be testable experimentally by measurements of fraction helical conformation of individual residues in oligopeptides of appropriate sequence. The relevance of these results in regards to the frequent occurrence of proline-containing helices in certain membrane proteins is discussed.  相似文献   

13.
We demonstrate a calculated alpha-helix peptide folding energy landscape which accurately simulates the first experimentally measured alpha-helix melting energy landscape. We examine a 21-amino acid, mainly polyalanine peptide and calculate the free energy along the Psi Ramachandran angle secondary folding coordinate. The experimental free energy landscape was determined using UV resonance Raman spectroscopy. The relative free energy values are very close as are the equilibrium peptide conformations. We find 2.3 kcal/mol activation barriers between the alpha-helix-like and PPII-like basins. We also find that the alpha-helix-like conformations are quite defective and the alpha-helix-like structure dynamically samples 310-helix and pi-bulges.  相似文献   

14.
An explanation in terms of conformational energies is provided for the observed nearly exclusive preference of the beta alpha beta structure for forming a right-handed, rather than a left-handed, crossover connection. Conformational energy computations have been carried out on a model beta alpha beta structure, consisting of two six-residue Val beta-strands and of a 12-residue Ala alpha-helix, connected by two flexible four-residue Ala links to the strands. The energy of the most favorable right-handed crossover is 15.51 kcal/mol lower than that of the corresponding left-handed cross-over. The right-handed crossover is a strain-free structure. Its energy of stabilization arises largely from the interactions of the two beta-strands with one another and with the alpha-helix. On the other hand, the left-handed crossover is either disrupted after energy minimization or it remains conformationally strained, as indicated by an energetically unfavorable left twisting of the beta-sheet and by the presence of high-energy local residue conformations. In the energetically most favorable right-handed crossover, the right twisting of the beta-sheet and its manner of interacting with the alpha-helix are identical with those computed earlier for isolated beta-sheets and for packed alpha/beta structures. This result supports a proposed principle that it is possible to account for the main features of frequently occurring structural arrangements in globular proteins in terms of the properties of their component structural elements.  相似文献   

15.
Molecular dynamics simulations of bee venom apamin, and an analogue having an Asn to Ala substitution at residue 2 (apamin-N2A), were analyzed to explore the contribution of hydrogen bonds involving Asn2 to local (beta-turn residues N2, C3, K4, A5) and global stability. The wild-type peptide retained a stable conformation during 2.4 ns of simulation at 67 degrees C, with high beta-turn stability characterized by backbone-side chain hydrogen bonds involving beta-turn residues K4 and A5, with the N2 side chain amide carbonyl. The loss of stabilizing interactions involving the N2 side chain resulted in the loss of the beta-turn conformation in the apamin N2A simulations (27 or 67 degrees C). This loss of beta-turn stability propagates throughout the peptide structure, with destabilization of the C-terminal helix connected to the N-terminal region by two disulfide bonds. Backbone stability in a synthetic peptide analogue (apamin-N2A) was characterized by NMR and amide hydrogen exchange measurements. Consistent with the simulations, loss of hydrogen bonds involving the N2 side chain resulted in destabilization of both the N-terminal beta-turn and the C-terminal helix. Amide exchange protection factors in the C-terminal helix were reduced by 9-11-fold in apamin N2A as compared with apamin, corresponding to free energy (deltaDeltaG(uf)) of around 1.5 kcal M(-1) at 20 degrees C. This is equivalent to the contribution of hydrogen bond interactions involving the N2 side chain to the stability of the beta-turn. Together with additional measures of exchange protection factors, the three main contributions to backbone stability in apamin that account for virtually the full thermodynamic stability of the peptide have been quantitated.  相似文献   

16.
We have recently designed a host-guest peptide system that allows us to quantitatively measure the energetics of interaction of viral fusion peptides with lipid bilayers. Here, we show that fusion peptides of influenza hemagglutinin reversibly associate with one another at membrane surfaces above critical surface concentrations, which range from one to five peptides per 1000 lipids in the systems that we investigated. It is further demonstrated by using circular dichroism and Fourier transform infrared spectroscopy that monomeric peptides insert into the bilayers in a predominantly alpha-helical conformation, whereas self-associated fusion peptides adopt predominantly antiparallel beta-sheet structures at the membrane surface. The two forms are readily interconvertible and the equilibrium between them is determined by the pH and ionic strength of the surrounding solution. Lowering the pH favors the monomeric alpha-helical conformation, whereas increasing the ionic strength shifts the equilibrium towards the membrane-associated beta-aggregates. The binding data are interpreted in terms of a cooperative binding model that yields free energies of insertion and free energies of self-association for each of the peptides studied at pH 7.4 and pH 5. At pH 5 and 35 mM ionic strength, the insertion energy of the 20 residue influenza hemagglutinin fusion peptide is -7.2 kcal/mol and the self-association energy is -1.9 kcal/mol. We propose that self-association of fusion peptides could be a major driving force for recruiting a small number of hemagglutinin trimers into a fusion site.  相似文献   

17.
The number of beta-turns in a representative set of 426 protein three-dimensional crystal structures selected from the recent Protein Data Bank has nearly doubled and the number of gamma-turns in a representative set of 320 proteins has increased over seven times since the previous analysis. Beta-turns (7153) and gamma-turns (911) extracted from these proteins were used to derive a revised set of type-dependent amino acid positional preferences and potentials. Compared with previous results, the preference for proline, methionine and tryptophan has increased and the preference for glutamine, valine, glutamic acid and alanine has decreased for beta-turns. Certain new amino acid preferences were observed for both turn types and individual amino acids showed turn-type dependent positional preferences. The rationale for new amino acid preferences are discussed in the light of hydrogen bonds and other interactions involving the turns. Where main-chain hydrogen bonds of the type NH(i + 3) --> CO(i) were not observed for some beta-turns, other main-chain hydrogen bonds or solvent interactions were observed that possibly stabilize such beta-turns. A number of unexpected isolated beta-turns with proline at i + 2 position were also observed. The NH(i + 2) --> CO(i) hydrogen bond was observed for almost all gamma-turns. Nearly 20% classic gamma-turns and 43% inverse gamma-turns are isolated turns.  相似文献   

18.
Jas GS  Kuczera K 《Biophysical journal》2004,87(6):3786-3798
We have performed experimental measurements and computer simulations of the equilibrium structure and folding of a 21-residue alpha-helical heteropeptide. Far ultraviolet circular dichroism spectroscopy is used to identify the presence of helical structure and to measure the thermal unfolding curve. The observed melting temperature is 296 K, with a folding enthalpy of -11.6 kcal/mol and entropy of -39.6 cal/(mol K). Our simulations involve 45 ns of replica-exchange molecular dynamics of the peptide, using eight replicas at temperatures between 280 and 450 K, and the program CHARMM with a continuum solvent model. In a 30-ns simulation started from a helical structure, conformational equilibrium at all temperatures was reached after 15 ns. This simulation was used to calculate the peptide melting curve, predicting a folding transition with a melting temperature in the 330-350 K range, enthalpy change of -10 kcal/mol, and entropy change of -30 cal/(mol K). The simulation results were also used to analyze the peptide structural fluctuations and the free-energy surface of helix unfolding. In a separate 15-ns replica-exchange molecular dynamics simulation started from the extended structure, the helical conformation was first attained after approximately 2.8 ns, and equilibrium was reached after 10 ns of simulation. These results showed a sequential folding process with a systematic increase in the number of hydrogen bonds until the helical state is reached, and confirmed that the alpha-helical state is the global free-energy minimum for the peptide at low temperatures.  相似文献   

19.
We present the initial findings of a theoretical study of hydrogen bond formation between two formamide molecules in water and in carbon tetrachloride. These systems were chosen as the simplest models for secondary structure formation in the polar environment near the protein surface and the apolar environment of the protein interior. We have employed thermodynamic simulation methods to obtain absolute binding free energies and free energy profiles for the formation of peptide hydrogen bonds in the two solvents. We find that the amide hydrogen bond is stable by 8.4 kcal/mol in CCl4, and by 0.3 kcal/mol in water. Our results indicate also that the hydrogen-bonded dimer is 2.2 kcal/mol more stable in water than it is in CCl4. We compare our results with those from experiment, and discuss their use in interpreting mechanisms of protein folding.  相似文献   

20.
I L Karle 《Biopolymers》1989,28(1):1-14
Preferred conformation and types of molecular folding are some of the topics that can be addressed by structure analysis using x-ray diffraction of single crystals. The conformations of small linear peptide molecules with 2-6 residues are affected by polarity of solvent, presence of water molecules, hydrogen bonding with neighboring molecules, and other packing forces. Larger peptides, both cyclic and linear, have many intramolecular hydrogen bonds, the effect of which outweighs any intermolecular attractions. Numerous polymorphs of decapeptides grown from a variety of solvents, with different cocrystallized solvents, show a constant conformation for each peptide. Large conformational changes occur, however, upon complexation with metal ions. A new form of free valinomycin grown from DMSO exhibits near three-fold symmetry with only three intramolecular hydrogen bonds. The peptide is in the form of a shallow bowl with a hydrophobic exterior. Near the bottom of the interior of the bowl are three carbonyl oxygens, spaced and directed so that they are in position to form three ligands to a K+, e.g., complexation can be completed by the three lobes containing the beta-bends closing over and encapsulating the K+ ion. In another example, free antamanide and the biologically inactive perhydro analogue, in which four phenyl groups become cyclic hexyl groups, have essentially the same folding of backbone and side chains. The conformation changes drastically upon complexation with Li+ or Na+. However, the metal ion complex of natural antamanide has a hydrophobic globlar form whereas the metal ion complex of the inactive perhydro analogue has a polar band around the middle. The structure results indicate that the antamanide molecule is in a complexed form during its biological activity. Single crystal x-ray diffraction structure analyses have identified the manner in which water molecules are essential to creating minipolar areas on apolar helices. Completely apolar peptides, such as membrane-active peptides, can acquire amphiphilic character by insertion of a water molecule into the helical backbone of Boc-Aib-Ala-Leu-Aib-Ala-Leu-Aib-Ala-Leu-Aib-OMe, for example. The C-terminal half assumes an alpha-helix conformation, whereas the N-terminal half is distorted by an insertion of a water molecule W(1) between N(Ala5) and O(Ala2), forming hydrogen bonds N(5)H...W(1) and W(1)...O(2). The distortion of the helix exposes C = O(Aib1) and C = O(Aib4) to the outside environment with the consequence of attracting additional water molecules. The leucyl side chains are on the other side of the molecule. Thus a helix with an apolar sequence can mimic an amphiphilic helix.  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