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 78 毫秒
1.
Some properties of α-helices of polyclycine and polyalanine, up to the decapeptide, were investigated by ab initio molecular-orbital calculations. These helices were found to be unstable relative to the corresponding “fully extended chain” conformation. The electric field of helices of 8–10 residues is about 20% stronger than that of models built from noninteracting monomers, for example. This is a result of cooperativity, which is essentially governed by the intramolecular hydrogen bonds. The cooperativity is manifest in all properties of the helices: relative stability, dipole moment, proton affinity, electrical potential. The electric potential of helices of three and four residues is such that their instability can be compensated for by a single charged group acting as an “initiator.” The computed proton affinity of the (Ala)8 α-helix is about 45 kcal/mol larger than that of formamide, which confirms that long helices may be protonated at the carboxyl end in solution.  相似文献   

2.
Many of the bilayer spanning segments of membrane transport proteins contain proline residues, and most of them are believed to occur in α-helical form. A proline residue in the middle of an α-helix is known to produce a bend in the helix, and recent studies have focused on characterizing such a bend at atomic level. In the present case, molecular dynamics (MD) studies are carried out on helix F model of bacteriorhodopsin (BR) Ace-(Ala)7-Trp-(Ala)2-Tyr-Pro-(Ala)2-Trp-(Ala)8-NHMe and compared with Ace-(Ala)7-Trp-(Ala)2-Tyr-(Ala)3-Trp-(Ala)8-NHMe in which the proline is replaced by alanine. The bend in the helix is characterized by structural parameters such as kink angle (α), wobble angle (θ), virtual torsion angle (ρ), and the hydrogen bond distance d (Op?3 … Np+1). The average values and the flexibility involved in these parameters are evaluated. The correlation among the bend related parameters are estimated. The equilibrium side chain orientations of tryptophan and tyrosine residues are discussed and compared with those found in the recently proposed model of bacteriorhodopsin. Finally, a detailed characterization of the bend in terms of secondary structures such as αI, αII and goniometric helices are discussed, which can be useful in the interpretation of the experimental results on the secondary structures of membrane proteins involving the proline residue. © 1993 Wiley-Liss, Inc.  相似文献   

3.
Jon Applequist 《Biopolymers》1981,20(2):387-397
Circular dichroic spectra and oscillator strengths of the π-π transition near 190 nm are calculated for helical (Gly)6 and (Ala)6 at 30° intervals of the backbone torsion angles (?,ψ) over the range -180° ≤ ? ≤ -60°, ?60° ≤ ψ ≤ 180°, using the partially dispersive normal mode treatment of the dipole interaction model. Polarizabilities of atoms and the NC′O group are those determined semiempirically in previous studies. Calculations for (Ala)6 at (?,ψ) angles corresponding to the α-helix, the poly(Pro) II helix, a collagen single helix, a poly-(MeAla) helix, and single β-helices are found to agree well with most of the available experimental data.  相似文献   

4.
The packing of α-helices and β-sheets in six αβ proteins (e.g. flavodoxin) has been analysed. The results provide the basis for a computer algorithm to predict the tertiary structure of an αβ protein from its amino acid sequence and actual assignment of secondary structure.The packing of an individual α-helix against a β-sheet generally involves two adjacent ± 4 rows of non-polar residues on the α-helix at the positions i, i + 4, i + 8, i + 1, i + 5, i + 9. The pattern of interacting β-sheet residues results from the twisted nature of the sheet surface and the attendant rotation of the side-chains. At a more detailed level, four of the α-helical residues (i + 1, i + 4, i + 5 and i + 8) form a diamond that surrounds one particular β-sheet residue, generally isoleucine, leucine or valine. In general, the α-helix sits 10 Å above the sheet and lies parallel to the strand direction.The prediction follows a combinational approach. First, a list of possible β-sheet structures (106 to 1014) is constructed by the generation of all β-sheet topologies and β-strand alignments. This list is reduced by constraints on topology and the location of non-polar residues to mediate the sheet/helix packing, and then rank-ordered on the extent of hydrogen bonding. This algorithm was uniformly applied to 16 αβ domains in 13 proteins. For every structure, one member of the reduced list was close to the crystal structure; the root-mean-square deviation between equivalenced Cα atoms averaged 5.6 Å for 100 residues. For the αβ proteins with pure parallel β-sheets, the total number of structures comparable to or better than the native in terms of hydrogen bonds was between 1 and 148. For proteins with mixed β-sheets, the worst case is glyceraldehyde-3-phosphate dehydrogenase, where as many as 3800 structures would have to be sampled. The evolutionary significance of these results as well as the potential use of a combinatorial approach to the protein folding problem are discussed.  相似文献   

5.
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   

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.
Proline-induced constraints in alpha-helices   总被引:9,自引:0,他引:9  
L Piela  G Némethy  H A Scheraga 《Biopolymers》1987,26(9):1587-1600
The disrupting effect of a prolyl residue on an α-helix has been analyzed by means of conformational energy computations. In the preferred, nearly α-helical conformations of Ac-Ala4-Pro-NHMe and of Ac-Ala7-Pro-Ala7-NHMe, only the residue preceding Pro is not α-helical, while all other residues can occur in the α-helical A conformation; i.e., it is sufficient to introduce a conformational change of only one residue in order to accommodate proline in a distorted α-helix. Other low-energy conformations exist in which the conformational state of three residues preceding proline is altered considerably; on the other hand, another conformation in which these three residues retain the near-α-helical A-conformational state (with up to 26° changes of their dihedral angles ? and ψ, and a 48° change in one ω from those of the ideal α-helix) has a considerably higher energy. These conclusions are not altered by the substitution of other residues in the place of the Ala preceding Pro. The conformations of the peptide chain next to prolyl residues in or near an α-helix have been analyzed in 58 proteins of known structure, based on published atomic coordinates. Of 331 α-helices, 61 have a Pro at or next to their N-terminus, 21 have a Pro next to their C-terminus, and 30 contain a Pro inside the helix. Of the latter, 16 correspond to a break in the helix, 9 are located inside distorted first turns of the helix, and 5 are parts of irregular helices. Thus, the reported occurrence of prolyl residues next to or inside observed α-helices in proteins is consistent with the computed steric and energetic requirements of prolyl peptides.  相似文献   

8.
Statistics are collected and analyzed for the possibility of hydrogen bonding in the secondary structures of globular proteins, based on geometric criteria. Double and bifurcated bonds are considered as pairs of admissible H-bonds with two proton donors or two proton acceptors, respectively. Most of such bonds belong to peptide groups in α-helices, with O i …N i + 3 nearly as frequent as O i …N i + 4; in contrast, most of the 3/10-helical segments are too short to have any. Alternating double and bifurcated bonds in α-helices form an apparently cooperative network structure. A typical α-helical segment perhaps carries two stretches of the H-bond network broken in the middle. The constituent H-bonds are nonlinear: the hydrogen atom is off the straight line connecting the proton donor and proton acceptor atoms. This deflection is larger for H i + 3 vs. bond line O i −N i + 3 than for H i + 4 vs. O i −N i + 4, and though the two kinds of bond have about the same length (exceeding those typical of low-molecular compounds), O i …N i + 4 must be stronger than O i …N i + 3. Double/bifurcated bonds are also not coplanar, i.e., hydrogen atoms are beyond the N…O…N (or O…N…O) plane. The text was submitted by the authors in English.  相似文献   

9.
Protein secondary structure elements are arranged in distinct structural motifs such as four-α-helix bundle, 8α/8β TIM-barrel, Rossmann dinucleotide binding fold, assembly of a helical rod. Each structural motif is characterized by a particular type of helix-helix interactions. A unique pattern of contacts is formed by interacting helices of the structural motif. In each type of fold, edges of the helix surface, which participate in the formation of helix-helix contacts with preceding and following helices, differ. This work shows that circular arrangements of the four, eight, and sixteen α-helices, which are found in the four-α-helical motif, TIM-barrel 8α/8β fold, and helical rod of 16.3¯ helices per turn correspondingly, can be associated with the mutual positioning of the edges of the helix surfaces. Edges (i, i+1)−(i+1, i+2) of the helix surface are central for the interhelical contacts in a four-α-helix bundle. Edges (i, i+1)−(i+2, i+3) are involved in the assembly of four-α-helix subunits into helical rod of a tobacco mosaic virus and a three-helix fragment of a Rossmann fold. In 8α/8β TIM-barrel fold, edges (i, i+1)−(i+5, i+6) are involved in the octagon arrangement. Approximation of a cross section of each motif with a polygon (n-gon, n=4, 8, 16) shows that a good correlation exists between polygon interior angles and angles formed by the edges of helix surfaces.  相似文献   

10.
Leader DP  Milner-White EJ 《Proteins》2011,79(3):1010-1019
We prepared a set of about 2000 α-helices from a relational database of high-resolution three-dimensional structures of globular proteins, and identified additional main chain i ← i+3 hydrogen bonds at the ends of the helices (i.e., where the hydrogen bonding potential is not fulfilled by canonical i ← i+4 hydrogen bonds). About one-third of α-helices have such additional hydrogen bonds at the N-terminus, and more than half do so at the C-terminus. Although many of these additional hydrogen bonds at the C-terminus are associated with Schellman loops, the majority are not. We compared the dihedral angles at the termini of α-helices having or lacking the additional hydrogen bonds. Significant differences were found, especially at the C-terminus, where the dihedral angles at positions C2 and C1 in the absence of additional hydrogen bonds deviate substantially from those occurring within the α-helix. Using a novel approach we show how the structure of the C-terminus of the α-helix can emerge from that of constituent overlapping α-turns and β-turns, which individually show a variation in dihedral angles at different positions. We have also considered the direction of propagation of the α-helix using this approach. If one assumes that helices start as a single α-turn and grow by successive addition of further α-turns, the paths for growth in the N → C and C → N directions differ in a way that suggests that extension in the C → N direction is favored.  相似文献   

11.
The concept of bends or chain reversals [nonhelical dipeptide sequences in which the distance R3 (i,i+3) between the Cα atoms of residues i and i+3 is ≦ 7.0 Å] has been extended to define double bends as tripeptide sequences, not in an α-helix, in which two successive distances R3(i,i+3) and R3 (i+1, i+4) are both ≦7.0 Å, with analogous definitions for higher-order multiple bends. A sample of 23 proteins, consisting of 4050 residues, contains 235 single, 58 double, and 11 higher-order multiple bends. Multiple bends may occur as combinations of the “standard” type I, II, and III chain reversals (as well as their mirror images), but usually they require distortions from these well-defined conformations. The frequency of occurrence of amino acids often differs significantly between single and multiple bends. The probability distribution of R3 distances does not differ in single and multiple bends. However, R4 (the distance between the Cα atoms of residues i and i+4) in multiple bends is generally shorter than in tripeptide sequences containing single bends. The value of R4 in many multiple bends is near those for α-helices. In some other multiple bends, R4 is even shorter, indicating that these structures are very compact. The signs of the dihedral angles about the virtual bonds connecting Cα atoms and the values of curvature and torsion, as defined by means of differential geometry, indicate that there is a preference for single and multiple bends to be right-handed (like an α-helical sequence, for example) and that there is a strong tendency to conserve the handedness in both single-bend components of many multiple bends. These often have a strong resemblance to distorted single turns of an α-helix and do not constitute chain reversals. Double bends, in which the signs of two successive virtual-bond dihedral angles differ, have conformations that are very different from an α-helix. They act as chain reversals occuring over three residues. These chain reversals have not been described previously. Multiple bends may play an important role in protein folding because they occur fairly frequently in proteins and cause major changes in the direction of the polypeptide chain.  相似文献   

12.
Alzheimer’s disease (AD) is a progressive neurodegenerative disorder that is characterized by loss of intellectual functioning of brain and memory loss. According to amyloid cascade hypothesis, aggregation of amyloid-β42 (Aβ42) peptide can generate toxic oligomers and their accumulation in the brain is responsible for the onset of AD. In spite of carrying out a large number of experimental studies on inhibition of Aβ42 aggregation by small molecules, the detailed inhibitory mechanism remains elusive. In the present study, comparable molecular dynamics (MD) simulations were performed to elucidate the inhibitory mechanism of a sulfonamide inhibitor C1 (2,5-dichloro-N-(4-piperidinophenyl)-3-thiophenesulfonamide), reported for its in vitro and in vivo anti-aggregation activity against Aβ42. MD simulations reveal that C1 stabilizes native α-helix conformation of Aβ42 by interacting with key residues in the central helix region (13–26) with hydrogen bonds and ππ interactions. C1 lowers the solvent-accessible surface area of the central hydrophobic core (CHC), KLVFF (16–20), that confirms burial of hydrophobic residues leading to the dominance of helical conformation in the CHC region. The binding free energy analysis with MM–PBSA demonstrates that Ala2, Phe4, Tyr10, Gln15, Lys16, Leu17, Val18, Phe19, Phe20, Glu22, and Met35 contribute maximum to binding free energy (?43.1 kcal/mol) between C1 and Aβ42 monomer. Overall, MD simulations reveal that C1 inhibits Aβ42 aggregation by stabilizing native helical conformation and inhibiting the formation of aggregation-prone β-sheet conformation. The present results will shed light on the underlying inhibitory mechanism of small molecules that show potential in vitro anti-aggregation activity against Aβ42.  相似文献   

13.
An apolar synthetic octapeptide, Boc-(Ala-Aib)4-OMe, was crystallized in the triclinic space group P1 with cell dimensions a = 11.558 Å, b = 11.643 Å, c = 9.650 Å, α = 120.220°, β = 107.000°, γ = 90.430°, V = 1055.889 Å3, Z = 1, C34H60O11N8·H2O. The calculated crystal density was 1.217 g/cm3 and the absorption coefficient ? was 6.1. All the intrahelical hydrogen bonds are of the 310 type, but the torsion angles, ? and ψ, of Ala(5) and Ala(7) deviate from the standard values. The distortion of the 310-helix at the C-terminal half is due to accommodation of the bulky Boc group of an adjacent peptide in the nacking. A water molecule is held between the N-terminal of one peptide and the C-terminal of the other. The oxygen atom of water forms hydrogen bonds with N (1) -H and N (2) -H, which are not involved in the intrahelical hydrogen bonds. The hydrogen atoms of water also formed hydrogen bonds with carbonyl oxygens of the adjacent peptide molecule. On the other hand, 1H-nmr analysis revealed that the octapeptide took an α-helical structure in a CD3CN solution. The longer peptides, Boc-(Ala-Aib)6-OMe and Boc-(Ala-Aib)8-OMe, were also shown to take an α-helical structure in a CD3CN solution. An α-helical conformation of the hexadecapeptide in the solid state was suggested by x-ray analysis of the crystalline structure. Thus, the critical length for transition from the 310- to α-helix of Boc-(Ala-Aib)n-OMe is 8. © 1993 John Wiley & Sons, Inc.  相似文献   

14.
《Biophysical chemistry》1986,25(2):201-213
Interactions between B-DNA and homopolymeric α-helices of glycine, alanine, serine, asparagine and aspartic acid have been studied theoretically. The complexation energy has been minimised taking into account the interactions between DNA and the polypeptides as well as the internal energy of the α-helix and the interaction energy of counterions with the complex. The results obtained indicate the important role of strong hydrogen bonds between the peptide side chains and nucleic acid phosphate groups, these bonds being much stronger than specific interactions with the base-pairs. The formation of these structural bonds depends on the size of the α-helix, which in turn determines whether bridging across the major groove is possible. The steric role of the methyl group of thymine in orienting the peptide helix and the role of DNA screening cations in complex stabilization are also significant.  相似文献   

15.
Conformations of Gly(n)H+ and Ala(n)H+ peptides in the gas phase   总被引:1,自引:0,他引:1       下载免费PDF全文
High-resolution ion mobility measurements and molecular dynamics simulations have been used to probe the conformations of protonated polyglycine and polyalanine (Gly(n)H and Ala(n)H+, n = 3-20) in the gas phase. The measured collision integrals for both the polyglycine and the polyalanine peptides are consistent with a self-solvated globule conformation, where the peptide chain wraps around and solvates the charge located on the terminal amine. The conformations of the small peptides are governed entirely by self-solvation, whereas the larger ones have additional backbone hydrogen bonds. Helical conformations, which are stable for neutral Alan peptides, were not observed in the experiments. Molecular dynamics simulations for Ala(n)H+ peptides suggest that the charge destabilizes the helix, although several of the low energy conformations found in the simulations for the larger Ala(n)H+ peptides have small helical regions.  相似文献   

16.
The conformational transitions of synthetic basic polytripeptides (Lys-Leu-Gly)n, (A2bu-Leu-Gly)n, (Lys-Leu-Ala)n, and (A2bu-Leu-Ala)n induced by high salt concentrations and elevated pH were investigated by CD, ir, and 1H-nmr spectroscopy, sedimentation analysis, viscometry, and light scattering. Sheet aggregates of chains in a conformation similar to the polyglycine II (polyproline II) helix, bound together by hydrogen bonds, are the most probable form of (Lys-Leu-Gly)n and also, partly, of (A2bu-Leu-Gly)n in a high-pH or high-salt solutions. The conformation (Lys-Leu-Ala)n, in a low-salt concentration, is an α-helix. Since (A2bu-Leu-Ala)n is disordered under similar conditions, it appears that this α-helix is stabilized by hydrophobic interactions between Lys and Leu side chains. In a high concentration of water structure-making ions, CD data for (Lys-Leu-Ala)n indicate distortion of the α-helix, with a parallel increase in the average molecular weight corresponding to trimer formation. Hydrodynamic data are consistent with a model of bundles of three closely touching spherocylinders. (A2bu-Leu-Ala)n shows a limited tendency to α-helix formation.  相似文献   

17.
The structure of the peptide Boc-Val-Ala-Leu-Aib-Val-Ala-Leu-OMe has been determined in crystals obtained from a dimethylsulfoxide–isopropanol mixture. Crystal parameters are as follows: C38H69N7O10 · H2O · 2C3H7OH, space group P21, a = 10.350 (2) Å, b = 26.084 (4) Å, c = 10.395(2) Å, β = 96.87(12), Z = 2, R = 8.7% for 2686 reflections observed > 3.0 σ (F). A single 5 → 1 hydrogen bond is observed at the N-terminus, while two 4 → 1 hydrogen bonds characteristic of a 310-helix are seen in the central segment. The C-terminus residues, Ala(6) and Leu(7) are expended, while Val(5) is considerably distorted from a helical conformation. Two isopropanol molecules make hydrogen bonds to the C-terminal segment, while a water molecule interacts with the N-terminus. The structure is in contrast to that obtained for the same peptide in crystals from methanol-water [ I. L. Karle, J. L. Flippen-Anderson, K. Uma, and P. Balaram (1990) Proteins: Structure, Function and Genetics, Vol. 7, pp. 62–73] in which two independent molecules reveal an almost perfect α-helix and a helix penetrated by a water molecule. A comparison of the three structures provides a snapshot of the progressive effects of solvation leading to helix unwinding. The fragility of the heptapeptide helix in solution is demonstrated by nmr studies in CDC13 and (CD3)2SO. A helical conformation is supported in the apolar solvent CDCl3, whereas almost complete unfolding is observed in the strongly solvating medium (CD3)2SO. © 1993 John Wiley & Sons, Inc.  相似文献   

18.
We have recorded high-resolution 13C-nmr spectra of collagen fibrils in the solid state by the cross-polarization–magic-angle-spinning(CP–MAS)method and analyzed the spectra with reference to those of collagenlike polypeptides. We used two kinds of model polypeptides to obtain reference 13C chemical shifts of major amino acid residues of collagen (Gly, Pro, Ala, and Hyp): the 31-helical polypeptides [(Gly)nII, (Pro)nII, (Hyp)n, and (Ala? Gly? Gly)nII], and the triple-helical polypeptides [(Pro? Gly? Pro)n and (Pro? Ala? Gly)n]. Examination of the 13C chemical shifts of these polypeptides, together with our previous data, showed that the 13C chemical shifts of individual amino acid residues are the same, within experimental error (±0.5 ppm), among different polypeptides with different primary sequences, if the conformations are the same. We found that the 13C chemical shifts of Ala residues of the 31-helical (Ala? Gly? Gly)n and triple-helical (Pro? Ala? Gly)n are significantly displaced, compared with those of the α-helix, β-sheet, and silk I form, and can be utilized as excellent probes to examine conformational features of collagen-like polypeptides. Further, the 13C chemical shifts of Gly and Pro residues in the triple-helical polypeptides are substantially displaced from those found in (Gly)nII and (Pro)nII of the 31-helix, reflecting further conformational change from the 31-helix to the supercoiled triple helix. In particular, the 13C chemical shifts of Gly C ? O carbons of the triple-helical polypeptides are substantially displaced upfield (4.1–5.1 ppm), with respect to those of the 31-helical polypeptides. These displacements are interpreted by that Gly C ? O of the former is not involved in NH …? O ? C hydrogen bonds, while this carbon of the latter is linked by these kinds of hydrogen bonds. On the basis of these 13C chemical shifts, as reference data for the collagenlike structure, we were able to assign the 13C-nmr peaks of Gly, Ala, Pro, and Hyp residues of collagen fibrils, which are in good agreement with the values expected from the model polypeptides mentioned above. We also discuss a plausible conformational change of collagen fibrils during denaturation.  相似文献   

19.
α-Aminoisobutyric acid (Aib) is a helicogenic α,α-dimethyl amino acid found in channel-forming peptaibols such as alamethicin. Possible effects of Aib on helix–helix packing are analyzed. Simulated annealing via restrained molecular dynamics is used to generate ensembles of approximately parallel helix dimers. Analysis of variations in geometrical and energetic parameters within ensembles defines how tightly a pair of helices interact. Simple hydrophobic helix dimers are compared: Ala20, Leu20, Aib20, and P20, the latter a simple channel-forming peptide [G. Menestrina, K. P. Voges, G, Jung, and G. Boheim (1986) Journal of Membrane Biology, Vol. 93, pp. 111–132]. Ala20 and Leu20 dimers exhibit well-defined ridges-in-grooves packing with helix crossing angles (Ω) of the order of +20°. Aib20 α-helix dimers are much more loosely packed, as evidenced by a wide range of Ω values and small helix-helix interaction energies. However, when in a 310 conformation Aib20 helices pack in three well-defined parallel modes, with Ω ca. ?15°, +5°, and 10°. Comparison of helix–helix interaction energies suggests that dimerization may favor the 310 conformation. P20, with 8 Aib residues, also shows looser packing of α-helices. The results of these studies of hydrophobic helix dimers are analyzed in the context of the ridges-in-grooves packing model. Simulations are extended to dimers of alamethicin, and of an alamethicin derivative in which all Aib residues are replaced by Leu. This substitution has little effect on helix–helix packing. Rather, such interactions appear to be sensitive to interactions between polar side chains. Overall, the results suggest that Aib may modulate the packing of simple hydrophobic helices, in favor of looser interactions. For more complex amphipathic helices, interactions between polar side chains may be more critical. © 1995 John Wiley & Sons, Inc.  相似文献   

20.
t-Buthyoxycarbonyl-L -alanyl-α-aminiosobutyryl-L -alanyl-α-aminoisobutyryl-α-aminoisobutyric acid methyl ester (t-Boc-L -Ala-Aib-L -Ala-Aib-Aib-OMe), C24H43N5O8, an end-protected pentapeptide with a sequence corresponding to the 6th through the 10th residues in suzukacillin, crystallizes in the orthorhombic space group P212121 with a = 11.671, b = 14.534, c = 17.906 Å and z = 4. The molecule exists as a right-handed 310-helix with a pitch of 6.026 Å. The helix is stabilized by three 4 → 1 hydrogen bonds with the NH groups of Ala(3), Aib(4), and Aib(5) hydrogen bonding to the carbonyl oxygens of t-Boc, Ala(1), and Aib(2), respectively. The helical molecules arrange themselves in a head-to-tail fashion along the a direction in such a way that the NH groups of Ala(1) and Aib(2) hydrogen bond to the carbonyl oxygens of Aib(4) and Aib(5), respectively, of a translationally related molecule. The helical columns thus formed close-pack nearly hexagonally to form the crystal.  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