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.
Models for the 3(10)-helix/coil and pi-helix/coil equilibria have been derived. The theory is based on classifying residues into helical or nonhelical (coil) conformations. Statistical weights are assigned to residues in a helical conformation with an associated helical hydrogen bond, a helical conformation with no hydrogen bond, an N-cap position, a C-cap position, or the reference coil conformation. The models for alpha-helix formation and 3(10)-helix formation have also been combined to describe a three-state equilibrium in which alpha-helical, 3(10)-helical, and coil conformations are populated. The results are compared with the modified Lifson-Roig theory for the alpha-helix/coil equilibrium. The comparison accounts for the experimental observations that 3(10)-helices tend to be short and pi-helices are not favored for any length. This work may provide a framework for quantitatively rationalizing experimental work on isolated 3(10)-helices and mixed 3(10)-/alpha-helices.  相似文献   

2.
Schievano E  Pagano K  Mammi S  Peggion E 《Biopolymers》2005,80(2-3):294-302
Aib-rich side-chain lactam-bridged oligomers Ac-(Glu-Aib-Aib-Lys)n-Ala-OH with n = 1,2,3 were designed and synthesized as putative models of the 3(10)-helix. The lactam bridge between the side chains of L-Glu and L-Lys in (i)--(i + 3) positions was introduced in order to enhance the structural preference toward the right-handed 3(10)-helix. The conformational properties of the three peptides were studied in trifluoroethanol (TFE) solution by CD, NMR, and computer simulations. The structural information was derived mainly from the analysis of nuclear Overhauser effect spectroscopy spectra. The presence of alpha H(i)-HN(i + 2) and of alpha H(i)-HN(i + 3) connectivities and the absence of alpha H(i)-HN(i + 4) connectivities indicate that these peptides fold into a 3(10)-helix rather than into an alpha-helix. Based on these conformational features, stereospecific assignment of the Aib methyl groups was possible. The results of such experiments and of the subsequent distance geometry and restrained molecular dynamics simulations reveal a marked preference of these peptides for 3(10)-helix. The CD spectra of these peptides indicate that the helix content increases upon chain elongation. The CD spectrum of the trimer is characterized by a negative band at 200 nm and by a weak positive band around 220 nm. The CD spectrum in TFE is different from that observed in aqueous solution in the presence of SDS micelles, reported in our previous work, and from those reported by a different research group for 3(10)-helical peptides. A possible reason for these differences could rest in the presence of different equilibria of the conformer populations of the various peptides in different solvent systems.  相似文献   

3.
The energy landscape of a peptide [Ace-Lys-Gln-Cys-Arg-Glu-Arg-Ala-Nme] in explicit water was studied with a multicanonical molecular dynamics simulation, and the AMBER parm96 force field was used for the energy calculation. The peptide was taken from the recognition helix of the DNA-binding protein, c-MYB: A rugged energy landscape was obtained, in which the random-coil conformations were dominant at room temperature. The CD spectra of the synthesized peptide revealed that it is in the random state at room temperature. However, the 300 K canonical ensemble, Q(300K), contained alpha-helix, 3(10)-helix, beta-turn, and beta-hairpin structures with small but notable probabilities of existence. The complete alpha-helix, imperfect alpha-helix, and random-coil conformations were separated from one another in the conformational space. This means that the peptide must overcome energy barriers to form the alpha-helix. The overcoming process may correspond to the hydrogen-bond rearrangements from peptide-water to peptide-peptide interactions. The beta-turn, imperfect 3(10)-helix, and beta-hairpin structures, among which there are no energy barriers at 300 K, were embedded in the ensemble of the random-coil conformations. Two types of beta-hairpin with different beta-turn regions were observed in Q(300K). The two beta-hairpin structures may have different mechanisms for the beta-hairpin formation. The current study proposes a scheme that the random state of this peptide consists of both ordered and disordered conformations. In contrast, the energy landscape obtained from the parm94 force field was funnel like, in which the peptide formed the helical conformation at room temperature and random coil at high temperature.  相似文献   

4.
Structures of N-termini of helices in proteins.   总被引:8,自引:7,他引:1       下载免费PDF全文
We have surveyed 393 N-termini of alpha-helices and 156 N-termini of 3(10)-helices in 85 high resolution, non-homologous protein crystal structures for N-cap side-chain rotamer preferences, hydrogen bonding patterns, and solvent accessibilities. We find very strong rotamer preferences that are unique to N-cap sites. The following rules are generally observed for N-capping in alpha-helices: Thr and Ser N-cap side chains adopt the gauche - rotamer, hydrogen bond to the N3 NH and have psi restricted to 164 +/- 8 degrees. Asp and Asn N-cap side chains either adopt the gauche - rotamer and hydrogen bond to the N3 NH with psi = 172 +/- 10 degrees, or adopt the trans rotamer and hydrogen bond to both the N2 and N3 NH groups with psi = 1-7 +/- 19 degrees. With all other N-caps, the side chain is found in the gauche + rotamer so that the side chain does not interact unfavorably with the N-terminus by blocking solvation and psi is unrestricted. An i, i + 3 hydrogen bond from N3 NH to the N-cap backbone C = O in more likely to form at the N-terminus when an unfavorable N-cap is present. In the 3(10)-helix Asn and Asp remain favorable N-caps as they can hydrogen bond to the N2 NH while in the trans rotamer; in contrast, Ser and Thr are disfavored as their preferred hydrogen bonding partner (N3 NH) is inaccessible. This suggests that Ser is the optimum choice of N-cap when alpha-helix formation is to be encouraged while 3(10)-helix formation discouraged. The strong energetic and structural preferences found for N-caps, which differ greatly from positions within helix interiors, suggest that N-caps should be treated explicitly in any consideration of helical structure in peptides or proteins.  相似文献   

5.
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   

6.
G Basu  A Kuki 《Biopolymers》1992,32(1):61-71
The analysis of the factors that control the helical folding of Aib-rich peptides is extended to include sensitivity to sequence patterns, and in particular the presence of contiguous non-Aib alpha-mono-alkylated residues. The distinct hydrogen-bonding network of the 3(10)-helix, as contrasted with that of the competing alpha-helical structure, is explicitly incorporated into a theoretical model for the 3(10)-helix/alpha-helix equilibrium constant for a given peptide. Finite length effects and the "extra" intrahelical hydrogen bond of the 3(10) form are expressed naturally as a result of this loop analysis. This semiempirical model captures all the established features of existing empirical rules for helical conformational transitions in Aib-rich sequences, as well as the recently detected helical transition induced solely by sequence permutation.  相似文献   

7.
The difference in length between the more elongated peptide 3(10)-helix and the more compact alpha-helix is about 0.4 A/residue. This property makes the 3(10)-/alpha-helix reversible conversion very promising as a molecular switching tool between the N- and C-terminal functions of a peptide backbone. In this work, using homo-peptides of various main-chain length, all based on the strongly helicogenic, Calpha-tetrasubstituted alpha-amino acid Calpha-methyl-L-valine, we show that a well defined, solvent controlled, reversible 3(10)-/alpha-helix transition takes place even in a homo-oligomer as short as a terminally blocked hexapeptide. Homo-peptide sequences blocked as a urethane or an acetamide at the N-terminus and as a methyl ester or an N-alkyl amide at the C-terminus are all appropriate. The nature of the occurring helical species in the various solvents tested was assessed by electronic or vibrational circular dichroism.  相似文献   

8.
Pal L  Basu G  Chakrabarti P 《Proteins》2002,48(3):571-579
An analysis of the shortest 3(10)-helices, containing three helical residues and two flanking capping residues that participate in two consecutive i + 3 --> i hydrogen bonds, shows that not all helices belong to the classic 3(10)-helix, where the three central residues adopt the right-handed helical conformation (alpha(R)). Three variants identified are: 3L10-helix with all residues in the left-handed helical region (alpha(L)), 3EL10-helix where the first residue is in the extended region followed by two residues in the alpha(L) conformation, and its mirror-image, the 3E'R10-helix. In the context of these helices, as well as the equivalent variants of alpha-helices, the length dependence of the handedness of secondary structures in protein structure is discussed. There are considerable differences in the amino acid preferences at different positions in the various types of 3(10)-helices. Each type of 3(10)-helix can be thought to be made up of an extension of a particular type of beta-turn (made up of residues i to i + 3) such that the (i + 3)th residue assumes the same conformation as the preceding residue. Distinct residue preferences at i and i + 3 positions seem to decide whether a particular stretch of four residues will be a beta-turn or a 3(10)-helix in the folded structure.  相似文献   

9.
As an additional step toward the dissection of the factors responsible for the onset of 3(10)-helix vs alpha-helix in peptides, in this paper we describe the results of a three-dimensional (3D) structural analysis by x-ray diffraction of the N(alpha)-acylated heptapeptide alkylamide mBrBz-L-Iva-L-(alphaMe)Val-L-Abu-L-(alphaMe)Val-L-(alphaMe)Phe-L-(alphaMe)Val-L-Iva-NHMe characterized by a single (L-Abu3) C(alpha)-trisubstituted and six C(alpha)-tetrasubstituted alpha-amino acids. We find that in the crystal state this peptide is folded in a mixed helical structure with short elements of 3(10)-helix at either terminus and a central region of alpha-helix. This finding, taken together with the published NMR and x-ray diffraction data on the all C(alpha)-methylated parent sequence and its L-Val2 analog (also the latter heptapeptide has a single C(alpha)-trisubstituted alpha-amino acid) strongly supports the view that one C(alpha)-trisubstituted alpha-amino acid inserted near the N-terminus of an N(alpha)-acylated heptapeptide alkylamide sequence may be enough to switch a regular 3(10)-helix into an essentially alpha-helical conformation. As a corollary of this work, the x-ray diffraction structure of the N(alpha)-protected, C-terminal tetrapeptide alkylamide Z-L-(alphaMe)Val-L-(alphaMe)Phe-L-(alphaMe)Val-L-Iva-NHMe, also reported here, is clearly indicative of the preference of this fully C(alpha)-methylated, short peptide for the 3(10)-helix. As the same terminally blocked sequence is mixed 3(10)/alpha-helical in the L-Abu3 heptapeptide amide but regular 3(10)-helical in the tetrapeptide amide and in the parent heptapeptide amide, these results point to an evident plasticity even of a fully C(alpha)-methylated short peptide.  相似文献   

10.
We introduce here i, i + 3 and i, i + 4 side chain interactions into the modified Lifson-Roig helix-coil theory of Doig et al. (1994, Biochemistry 33:3396-3403). The helix/coil equilibrium is a function of initiation, propagation, capping, and side chain interaction parameters. If each of these parameters is known, the helix content of any isolated peptide can be predicted. The model considers every possible conformation of a peptide, is not limited to peptides with only a single helical segment, and has physically meaningful parameters. We apply the theory to measure the i, i + 4 interaction energies between Phe and Met side chains. Peptides with these residues spaced i, i + 4 are significantly more helical than controls where they are spaced i, i + 5. Application of the model yields delta G for the Phe-Met orientation to be -0.75 kcal.mol-1, whereas that for the Met-Phe orientation is -0.54 kcal.mol-1. These orientational preferences can be explained, in part, by rotamer preferences for the interacting side chains. We place Phe-Met i, i + 4 at the N-terminus, the C-terminus, and in the center of the host peptide. The model quantitatively predicts the observed helix contents using a single parameter for the side chain-side chain interaction energy. This result indicates that the model works well even when the interaction is at different locations in the helix.  相似文献   

11.
A structural transition from a 3(10)-helix to an alpha-helix has been characterized at high resolution for an octapeptide segment located in 3 different sequences. Three synthetic peptides, decapeptide (A) Boc-Aib-Trp-(Leu-Aib-Ala)2-Phe-Aib-OMe, nonapeptide (B) Boc-Trp-(Leu-Aib-Ala)2-Phe-Aib-OMe, and octapeptide (C) Boc-(Leu-Aib-Ala)2-Phe-Aib-OMe, are completely helical in their respective crystals. At 0.9 A resolution, R factors for A, B, and C are 8.3%, 5.4%, and 7.3%, respectively. The octapeptide and nonapeptide form ideal 3(10)-helices with average torsional angles phi(N-C alpha) and psi(C alpha-C') of -57 degrees, -26 degrees C and -60 degrees, -27 degrees for B. The 10-residue peptide (A) begins as a 3(10)-helix and abruptly changes to an alpha-helix at carbonyl O(3), which is the acceptor for both a 4-->1 hydrogen bond with N(6)H and a 5-->1 hydrogen with N(7)H, even though the last 8 residues have the same sequence in all 3 peptides. The average phi, psi angles in the decapeptide are -58 degrees, -28 degrees for residues 1-3 and -63 degrees, -41 degrees for residues 4-10. The packing of helices in the crystals does not provide any obvious reason for the transition in helix type. Fourier transform infrared studies in the solid state also provide evidence for a 3(10)- to alpha-helix transition with the amide I band appearing at 1,656-1,657 cm-1 in the 9- and 10-residue peptides, whereas in shorter sequences the band is observed at 1,667 cm-1.  相似文献   

12.
Kunio Takeda 《Biopolymers》1985,24(4):683-694
Conformational changes of poly(L-ornithine) [(Orn)n] were studied in a sodium dodecyl sulfate (NaDodSO4) solution by CD. (Orn)n adopted an unstable and a stable helical structure below and above the NaDodSO4 concentration range where β-structure was favored, respectively. CD stopped-flow was used to monitor the transitions from coil to the unstable helix, from the helix to β-structure, and from coil to β-structure. Only the rate of the helix to β-structure transition was accelerated by an increase in NaDodSO4 concentration, whereas the rates of the others were independent of NaDodSO4 concentration. The fractions of coil, α-helix, and β-structure in each conformation of (Orn)n caused by NaDodSO4 were computed by simulating a mixed spectrum of typical CD spectra for these structures to the experimentally obtained spectrum. The contents of the unstable and stable helical structures were less than 50 and 73%, respectively.  相似文献   

13.
Local determinants of 3(10)-helix stabilization have been ascertained from the analysis of the crystal structure data base. We have clustered all 5-length substructures from 51 nonhomologous proteins into classes based on the conformational similarity of their backbone dihedral angles. Several clusters, derived from 3(10)-helices and multiple-turn conformations, had strong amino acid sequence patterns not evident among alpha-helices. Aspartate occurred over twice as frequently in the N-cap position of 3(10)-helices as in the N-cap position of alpha-helices. Unlike alpha-helices, 3(10)-helices had few C-termini ending in a left-handed alpha conformation; most 3(10) C-caps adopted an extended conformation. Differences in the distribution of hydrophobic residues among 3(10)- and alpha-helices were also apparent, producing amphipathic 3(10)-helices. Local interactions that stabilize 3(10)-helices can be inferred both from the strong amino acid preferences found for these short helices, as well as from the existence of substructures in which tertiary interactions replace consensus local interactions. Because the folding and unfolding of alpha-helices have been postulated to proceed through reverse-turn and 3(10)-helix intermediates, sequence differences between 3(10)- and alpha-helices can also lend insight into factors influencing alpha-helix initiation and propagation.  相似文献   

14.
Straight-chain, non-natural, nonpolar amino acids norleucine, norvaline, and alpha-amino-n-butyric acid at various spacings do not interact with themselves to stabilize helix formation in alanine-based peptides, but do interact with a Tyr spaced i, i + 4 to stabilize alanine helices, similar to the helix-stabilizing i, i + 4 Tyr-Leu and Tyr-Val interactions reported earlier (Padmanabhan S, Baldwin RL, 1994, J Mol Biol 241:706-713). Leu spaced i, i + 4 from another Leu is measurably helix-stabilizing relative to the corresponding i, i + 3 pair, but less so than for i, i + 4 Val-Leu, Ile-Leu, or Phe-Leu pairs (relative to the corresponding i, i + 3 pairs) when Leu is C-terminal to the other nonpolar amino acid. Our results indicate that limited side-chain flexibility in an alpha-helix strongly favors the interaction between 2 nonpolar residues to stabilize an isolated alpha-helix.  相似文献   

15.
The observation that short, linear alanine-based polypeptides form stable α-helices in aqueous solution has allowed the development of well-defined experimental systems with which to study the influence of amino acid sequence upon the stability of secondary structure. We have performed detailed conformational searches upon six alanine-based peptides in order to rationalize the observed variation in the α-helical stability in terms of side-chain-backbone and side-chain-side-chain interactions. Although a simple, gas-phase, potential model was used to obtain the conformational energies for these peptides, good agreement was obtained with experiment regarding their relative α-helical stabilities. Our calculations clearly indicate that valine, isoleucine, and phenylalanine residues should destabilize the α-helical conformation when included within alanine-based peptides because of energetically unfavorable side-chain-backbone interactions, which tend to result in the formation of regions of 310-helix. In the case of valine, the destabilization most probably arises from entropic effects as the isopropyl side chain can assume more orientations in the 310-helical form of the peptide. A detailed examination of very short-range interactions in these peptides has also indicated that an interaction, involving fewer than five consecutive residues, whose stabilizing effect reinforces that of the (i, i + 4) hydrogen bond may be the basis of the requirement for increased nucleation (σ) and propagation parameters (s) required by Zimm–Bragg theory to predict the α-helical content for compounds in this class of short peptides. Our calculations complement recent work using modified Zimm–Bragg and Lifson–Roig theories of the helix–coil transition, and are consistent with molecular dynamics simulations upon linear peptides in aqueous solution. © 1993 John Wiley & Sons, Inc.  相似文献   

16.
An x-ray crystallographic analysis was carried out for Boc-(Aib-DeltaZPhe)4-Aib-OMe (1: Boc = t-butoxycarbonyl; Aib = alpha-aminoisobutyric acid; DeltaZPhe = Z-alpha,beta-didehydrophenylalanine) to provide the precise conformational parameters of the octapeptide segment -(Aib-DeltaZPhe)4-. Peptide 1 adopted a typical 3(10)-helical conformation characterized by = +/-55.8 degrees (50 degrees -65 degrees), = +/-26.7 degrees (15 degrees -45 degrees), and = +/-179.5 degrees (168 degrees -188 degrees) for the average values of the -(Aib-DeltaZPhe)4- segment (the range of the eight values). The 3(10)-helix contains 3.1 residues per turn, being close to the "perfect 3(10)-helix" characterized by 3.0 residues per turn. NMR and Fourier transform infrared (FTIR) spectroscopy revealed that the 3(10)-helical conformation at the atomic resolution is essentially maintained in solution. Energy minimization of peptide 1 by semiempirical molecular orbital calculation converged to a 3(10)-helical conformation similar to the x-ray crystallographic 3(10)-helix. The preference for a 3(10)-helix in the -(Aib-DeltaZPhe)4- segment is ascribed to strong inducers of the 3(10)-helix inherent in Aib and DeltaZPhe residues-in particular, the Aib residues tend to stabilize a 3(10)-helix more effectively. Therefore, the -(Aib-DeltaZPhe)4- segment is useful to rationally design an optically inactive 3(10)-helical backbone, which will be of great importance to provide novel insights into noncovalent and covalent chiral interactions of a helical peptide with a chiral molecule.  相似文献   

17.
The kinetics of α-helix formation in polyalanine and polyglycine eicosamers (20-mers) were examined using torsional-coordinate molecular dynamics (MD). Of one hundred fifty-five MD experiments on extended (Ala)20 carried out for 0.5 ns each, 129 (83%) formed a persistent α-helix. In contrast, the extended state of (Gly)20 only formed a right-handed α-helix in two of the 20 MD experiments (10%), and these helices were not as long or as persistent as those of polyalanine. These simulations show helix formation to be a competition between the rates of (a) forming local hydrogen bonds (i.e. hydrogen bonds between any residue i and its i + 2, i + 3, i + 4, or i + 5th neighbor) and (b) forming nonlocal hydrogen bonds (HBs) between residues widely separated in sequence. Local HBs grow rapidly into an α-helix; but nonlocal HBs usually retard helix formation by “trapping” the polymer in irregular, “balled-up” structures. Most trajectories formed some nonlocal HBs, sometimes as many as eight. But, for (Ala)20, most of these eventually rearranged to form local HBs that lead to α-helices. A simple kinetic model describes the rate of converting nonlocal HBs into α-helices. Torsional-coordinate MD speeds folding by eliminating bond and angle degrees of freedom and reducing dynamical friction. Thus, the observed 210 ps half-life for helix formation is likely to be a lower bound on the real rate. However, we believe the sequential steps observed here mirror those of real systems. Proteins 33:343–357, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

18.
Biron Z  Khare S  Samson AO  Hayek Y  Naider F  Anglister J 《Biochemistry》2002,41(42):12687-12696
The peptide gp41(659-671) (ELLELDKWASLWN) comprises the entire epitope for one of the three known antibodies capable of neutralizing a broad spectrum of primary HIV-1 isolates and is the only such epitope that is sequential. Here we present the NMR structure of gp41(659-671) in water. This peptide forms a monomeric 3(10)-helix stabilized by i,i+3 side chain-side chain interactions favored by its primary sequence. In this conformation the peptide presents an exposed surface, which is mostly hydrophobic and consists of conserved HIV-1 residues. The presence of the 3(10)-helix is confirmed by its characteristic CD pattern. Studies of the 3(10)-helix have been hampered by the absence of a model peptide adopting this conformation. gp41(659-671) can serve as such a model to investigate the spectral characteristics of the 3(10)-helix, the factors that influence its stability, and the propensity of different amino acids to form a 3(10)-helix. The observation that the 3(10)-helical conformation is highly populated in the peptide gp41(659-671) indicates that the corresponding segment in the cognate protein is an autonomous folding unit. As such, it is very likely that the helical conformation is maintained in gp41 throughout the different tertiary structures of the envelope protein that form during the process of viral fusion. However, the exposure of the gp41(659-671) segment may vary, leading to changes in the reactivity of anti-gp41 antibodies in the different stages of viral fusion. Since gp41(659-671) is an autonomous folding unit, peptide immunogens consisting of the complete gp41(659-671) sequence are likely to induce antibodies highly cross-reactive with HIV-1.  相似文献   

19.
A series of short, amphipathic peptides incorporating 80% C(alpha),C(alpha)-disubstituted glycines has been prepared to investigate amphipathicity as a helix-stabilizing effect. The peptides were designed to adopt 3(10)- or alpha-helices based on amphipathic design of the primary sequence. Characterization by circular dichroism spectroscopy in various media (1 : 1 acetonitrile/water; 9 : 1 acetonitrile/water; 9 : 1 acetonitrile/TFE; 25 mM SDS micelles in water) indicates that the peptides selectively adopt their designed conformation in micellar environments. We speculate that steric effects from ith and ith + 3 residues interactions may destabilize the 3(10)-helix in peptides containing amino acids with large side-chains, as with 1-aminocyclohexane-1-carboxylic acid (Ac(6)c). This problem may be overcome by alternating large and small amino acids in the ith and ith + 3 residues, which are staggered in the 3(10)-helix.  相似文献   

20.
The conformational equilibrium between 3(10)- and alpha-helical structure has been studied via high-resolution NMR spectroscopy by Millhauser and coworkers using the MW peptide Ac-AMAAKAWAAKA AAARA-NH2. Their 750-MHz nuclear Overhauser effect spectroscopy (NOESY) spectra were interpreted to reflect appreciable populations of 3(10)-helix throughout the peptide, with the greatest contribution at the N and C termini. The presence of simultaneous alphaN(i,i + 2) and alphaN(i,i + 4) NOE cross-peaks was proposed to represent conformational averaging between 3(10)- and alpha-helical structures. In this study, we describe 25-nsec molecular dynamics simulations of the MW peptide at 298 K, using both an 8 A and a 10 A force-shifted nonbonded cutoff. The ensemble averages of both simulations are in reasonable agreement with the experimental helical content from circular dichroism (CD), the (3)J(HNalpha) coupling constants, and the 57 observed NOEs. Analysis of the structures from both simulations revealed very little formation of contiguous i --> i + 3 hydrogen bonds (3(10)-helix); however, there was a large population of bifurcated i --> i + 3 and i --> i + 4 alpha-helical hydrogen bonds. In addition, both simulations contained considerable populations of pi-helix (i --> i + 5 hydrogen bonds). Individual turns formed over residues 1-9, which we predict contribute to the intensities of the experimentally observed alphaN(i,i + 2) NOEs. Here we show how sampling of both folded and unfolded structures can provide a structural framework for deconvolution of the conformational contributions to experimental ensemble averages.  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