首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
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   

2.
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   

3.
This work demonstrates that all packing in α-helices can be simplified to repetitive patterns of a single motif: the knob-socket. Using the precision of Voronoi Polyhedra/Delauney Tessellations to identify contacts, the knob-socket is a four-residue tetrahedral motif: a knob residue on one α-helix packs into the three-residue socket on another α-helix. The principle of the knob-socket model relates the packing between levels of protein structure: the intra-helical packing arrangements within secondary structure that permit inter-helix tertiary packing interactions. Within an α-helix, the three-residue sockets arrange residues into a uniform packing lattice. Inter-helix packing results from a definable pattern of interdigitated knob-socket motifs between two α-helices. Furthermore, the knob-socket model classifies three types of sockets: (1) free, favoring only intra-helical packing; (2) filled, favoring inter-helical interactions; and (3) non, disfavoring α-helical structure. The amino acid propensities in these three socket classes essentially represent an amino acid code for structure in α-helical packing. Using this code, we used a novel yet straightforward approach for the design of α-helical structure to validate the knob-socket model. Unique sequences for three peptides were created to produce a predicted amount of α-helical structure: mostly helical, some helical, and no helix. These three peptides were synthesized, and helical content was assessed using CD spectroscopy. The measured α-helicity of each peptide was consistent with the expected predictions. These results and analysis demonstrate that the knob-socket motif functions as the basic unit of packing and presents an intuitive tool to decipher the rules governing packing in protein structure.  相似文献   

4.
Conformational energy computations were carried out on the packing of two identical collagenlike poly(tripeptide) triple helices in order to determine the energetics of favorable packing arrangements as a function of composition and chain length. The triple helices considered were [CH3CO-(Gly-Pro-Pro)nt-NHCH3]3 and [CH3CO-(Gly-Pro-Ala)nt-NHCH3]3, with nt = 3, 4, and 5. The packing arrangements were characterized in terms of their intermolecular energies and orientation angles Ω0 of the axes of the two triple helices. For short triple helices (nt = 3 or 4), many low-energy orientations, with a wide range of values of Ω0, can occur. When the triple helices are longer (nt = 5), the only low-energy packing arrangements of two poly(Gly-Pro-Pro) triple helices are those with a nearly parallel orientation of the two helix axes, with Ω0 ≈ ?10°. This result accounts for the observed parallel (rather than antiparallel) arrangement of collagen molecules in microfibril assembly and stands in contrast to the preferred antiparallel arrangement of a pair of α-helices. Since the preference for a parallel arrangement of these collagenlike triple helices is less pronounced in the case of poly(Gly-Pro-Ala), it appears that this preference is a consequence of the frequent presence of imino acids in position Y of the Gly-X-Y repeating triplet. In poly(Gly-Pro-Ala), most of the low-energy packing arrangements are parallel, but a few arrangements with low energies and high values of |Ω0| occur. These packing arrangements have a high energy, however, when Pro is substituted for Ala, and thus they are not accessible for collagen with natural amino (imino) acid sequences. The computations reported here account for some of the characteristic features of collagen packing in terms of the local interaction energies of a pair of triple helices.  相似文献   

5.
Abstract—Mutual arrangement, or packing, of α-helices in proteins depends on several factors, but, tight packing and the chemical nature of the polypeptide chain are the most important. This study shows, for the first time, that the torsion packing angles between axes of α-helices depend on their length. A database of helical pairs formed by two connected and juxtaposed α-helices has been compiled using the Protein Data Bank. These helical pairs have been subdivided into four types: (1) 10474 pairs formed by long helices; (2) 3665 pairs in which the first α-helix is long and the second is short; (3) 3648 pairs in which the first α‑helix is short and the second is long; 4) 1895 pairs in which both helices are short. Analysis of the database showed that most helical pairs in which both the helices are long form α-hairpins having interhelical packing angles of Ω ≈ 20°. Most helical pairs in which one α-helix is long and the other is short or both helices are short form αα-corners having orthogonal (Ω ≈ –70°…–90°) or slanted (Ω ≈ –50°) packing of α-helices. The possible reasons for this relationship between interhelical angles (Ω) and the length of α-helices are discussed. These results are of great importance in protein modeling and prediction since they enable the determination of the mutual arrangement of α-helices in protein molecules.  相似文献   

6.
Stabilized α-helices and nonpeptidic helix mimetics have emerged as powerful molecular scaffolds for the discovery of protein-protein interaction inhibitors. Protein-protein interactions often involve large contact areas, which are often difficult for small molecules to target with high specificity. The hypothesis behind the design of stabilized helices and helix mimetics is that these medium-sized molecules may pursue their targets with higher specificity because of a larger number of contacts. This protocol describes an optimized synthetic strategy for the preparation of stabilized α-helices that feature a carbon-carbon linkage in place of the characteristic N-terminal main-chain hydrogen bond of canonical helices. Formation of the carbon-carbon bond is enabled by a microwave-assisted ring-closing metathesis reaction between two terminal olefins on the peptide chain. The outlined strategy allows the synthesis and purification of a hydrogen bond surrogate (HBS) α-helix in ~ 1 week.  相似文献   

7.
A modification of the α-helix, termed the ω-helix, has four residues in one turn of a helix. We searched the ω-helix in proteins by the HELFIT program which determines the helical parameters—pitch, residues per turn, radius, and handedness—and p = rmsd/(N ? 1)1/2 estimating helical regularity, where “rmsd” is the root mean square deviation from the best fit helix and “N” is helix length. A total of 1,496 regular α-helices 6–9 residues long with p ≤ 0.10 Å were identified from 866 protein chains. The statistical analysis provides a strong evidence that the frequency distribution of helices versus n indicates the bimodality of typical α-helix and ω-helix. Sixty-two right handed ω-helices identified (7.2% of proteins) show non-planarity of the peptide groups. There is amino acid preference of Asp and Cys. These observations and analyses insist that the ω-helices occur really in proteins.  相似文献   

8.
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   

9.
General architecture of the alpha-helical globule   总被引:4,自引:0,他引:4  
A model is presented for the arrangement of alpha-helices in globular proteins. In the model, helices are placed on certain ribs of "quasi-spherical" polyhedra. The polyhedra are chosen so as to allow the close packing of helices around a hydrophobic core and to stress the collective interactions of the individual helices. The model predicts a small set of stable architectures for alpha-helices in globular proteins and describes the geometries of the helix packings. Some of the predicted helix arrangements have already been observed in known protein structures; others are new. An analysis of the three-dimensional structures of all proteins for which co-ordinates are available shows that the model closely approximates the arrangements and packing of helices actually observed. The average deviations of the real helix axes from those in the model polyhedra is +/- 20 degrees in orientation and +/- 2 A in position (1 A = 0.1 nm). We also show that for proteins that are not homologous, but whose helix arrangements are described by the same polyhedron, the root-mean-square difference in the position of the C alpha atoms in the helices is 1.6 to 3.0 A.  相似文献   

10.
A sequence analysis and comparison of transmembrane helices in bacteriorhodopsin (BR) and G protein-coupled receptors (GPCRs) is presented to identify potential regions of homology across protein families. The results show a common pattern of residues is conserved within the interhelical contact regions of BR that fit a knob-into-hole structural motif previously postulated for globular proteins and photosynthetic reaction centers. Based on an alignment of conserved prolines in transmembrane helices, it is inferred that analogous helix packing arrangements are possible in the rhodopsin-like GPCRs. Molecular models of GPCR helices V and VI indicate these interactions occur between aromatic and hydrophobic residues flanking the highly conserved prolines in these sequences. A similar packing arrangement is shown to occur in the X-ray structure of the melittin which also displays a unique pairing of proline-linked helices. The contact pattern identified is further applied to predict the packing of pairs of proline-containing helices in the pheromone-like and cAMP GPCRs. A potential role in stabilizing structure formation is also suggested for the contacts. The results and conclusions are supported by recent biophysical studies of zinc binding to kappa-opioid receptor mutants.  相似文献   

11.
The formation of α-helical assembly by complexing biologically active peptides with de novo designed protein is described. The de novo designed protein described here is a cystinelinked 4-helix bundle protein constructed with 80 amino acid residues and forms a hydrophobic core region surrounded by 4 helices in an aqueous solution. The biologically active peptides, such as melittin and human growth hormone releasing factor, contain the sequences that are able to form amphiphilic helices. These peptides alone do not form the α-helix structure in a diluted solution with low ion strength. But on mixing with the designed helix bundle protein, the peptides are strongly bound to the protein with the induction of α-helical structure in the biologically active peptides. The content of induced α-helix is in accord with that estimated from the amphiphilic sequence. The results mean that a novel architecture composed of α-helices is formed. Fluorescent and temperature-scanning measurement revealed that the α-helical assembly is constructed with hydrophobic interaction. Also, it is shown by means of fluorescence depolarization that the assembly has a compact globular form corresponding to 1 : 1 complex. © 1994 John Wiley & Sons, Inc.  相似文献   

12.
Apolipophorin III (apoLp-III) is an exchangeable apolipoprotein whose structure is represented as a bundle of five amphipathic alpha-helices. In order to study the properties of the helical domains of apolipophorin III, we designed and obtained five single-tryptophan mutants of Locusta migratoria apoLp-III. The proteins were studied by UV absorption spectroscopy, time-resolved and steady-state fluorescence spectroscopy, and circular dichroism. Fluorescence anisotropy, near-UV CD and solute fluorescence quenching studies indicate that the Trp residues in helices 1 (N-terminal) and 5 (C-terminal) have the highest conformational flexibility. These two residues also showed the highest degree of hydration. Trp residues in helices 3 and 4 display the lowest mobility, as assessed by fluorescence anisotropy and near UV CD. The Trp residue in helix 2 is protected from the solvent but shows high mobility. As inferred from the properties of the Trp residues, helices 1 and 5 appear to have the highest conformational flexibility. Helix 2 has an intermediate mobility, whereas helices 3 and 4 appear to constitute a highly ordered domain. From the configuration of the helices in the tertiary structure of the protein, we estimated the relative strength of the five interhelical interactions of apoLp-III. These interactions can be ordered according to their apparent stabilizing strengths as: helix 3-helix 4 > helix 2-helix 3 > helix 4-helix 1 approximately helix 2-helix 5 > helix 1-helix 5. A new model for the conformational change that is expected to occur upon binding of the apolipoprotein to lipid is proposed. This model is significantly different from the currently accepted model (Breiter, D. R., Kanost, M. R., Benning, M. M., Wesemberg, G., Law, J. H., Wells, M. A., Rayment, I., and Holden, M. (1991) Biochemistry 30, 603-608). The model presented here predicts that the relaxation of the tertiary structure and the concomitant exposure of the hydrophobic core take place through the disruption of the weak interhelical contacts between helices 1 and 5. To some extent, the weakness of the helix 1-helix 5 interaction would be due to the parallel arrangement of these helices.  相似文献   

13.
The determination of the crystal structure of the Ca2+-ATPase of sarcoplasmic reticulum (SR) in its Ca2+-bound [Nature 405 (2000) 647] and Ca2+-free forms [Nature 418 (2002) 605] gives the opportunity for an analysis of conformational changes on the Ca2+-ATPase and of helix-helix and helix-lipid interactions in the transmembrane (TM) region of the ATPase. The locations of the ends of the TM α-helices on the cytoplasmic side of the membrane are reasonably well defined by the location of Trp residues and by the location of Lys-262 that snorkels up to the surface. The locations of the lumenal ends of the helices are less clear. The position of Lys-972 on the lumenal side of helix M9 suggests that the hydrophobic thickness of the protein is only about 21 Å, rather than the normal 30 Å. The experimentally determined TM α-helices do not agree well with those predicted theoretically. Charged headgroups are required for strong interaction of lipids with the ATPase, consistent with the large number of charged residues located close to the lipid-water interface. Helix packing appears to be rather irregular. Packing of helices M8 and M10 is of the 3-4 ridges-into-grooves or knobs-into-holes types. Packing of helices M5 and M7 involves two Gly residues in M7 and one Gly residue in M5. Packing of the other helices generally involves just one or two residues on each helix at the crossing point. The irregular packing of the TM α-helices in the Ca2+-ATPase, combined with the diffuse structure of the ATPase on the lumenal side of the membrane, is suggested to lead to a relative low activation energy for changing the packing of the TM α-helices, with changes in TM α-helical packing being important in the process of transfer of Ca2+ ions across the membrane. The inhibitor thapsigargin binds in a cleft between TM α-helices M3, M5 and M7. It is suggested that this and other similar clefts provide binding sites for a variety of hydrophobic molecules affecting the activity of the Ca2+-ATPase.  相似文献   

14.
Structure prediction of membrane proteins could be constrained and thereby improved by introducing data of the observed molecular shape. We studied a coarse-grained molecular model that relied on residue-based dummy atoms to fold the transmembrane helices of a protein in the observed molecular shape. Based on the inter-residue potential, the α-helices were folded to contact each other in a simulated annealing protocol to search optimized conformation. Fitting the model into a three-dimensional volume was tested for proteins with known structures and resulted in a fairly reasonable arrangement of helices. In addition, the constraint to the packing transmembrane helix with the two-dimensional region was tested and found to work as a very similar folding guide. The obtained models nicely represented α-helices with the desired slight bend. Our structure prediction method for membrane proteins well demonstrated reasonable folding results using a low-resolution structural constraint introduced from recent cell-surface imaging techniques.  相似文献   

15.
《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   

16.
Murthy VL  Rose GD 《Biochemistry》2000,39(47):14365-14370
Although energetic and phylogenetic methods have been very successful for prediction of nucleic acid secondary structures, arrangement of these secondary structure elements into tertiary structure has remained a difficult problem. Here we explore the packing arrangements of DNA, RNA, and DNA/RNA hybrid molecules in crystals. In the conventional view, the highly charged double helix will be pushed toward isolation by favorable solvation effects; interactions with other like-charged stacks would be strongly disfavored. Contrary to this expectation, we find that most of the cases analyzed ( approximately 80%) exhibit specific, preferential packing between elements of secondary structure, which falls into three categories: (i) interlocking of major grooves of two helices, (ii) side-by-side parallel packing of helices, and (iii) placement of the ribose-phosphate backbone ridge of one helix into the major groove of another. The preponderance of parallel packing motifs is especially surprising. This category is expected to be maximally disfavored by charge repulsion. Instead, it comprises in excess of 50% of all packing interactions in crystals of A-form RNA and has also been observed in crystal structures of large RNA molecules. To explain this puzzle, we introduce a novel model for RNA folding. A simple calculation suggests that the entropy gained by a cloud of condensed cations surrounding the helices more than offsets the Coulombic repulsion of parallel arrangements. We propose that these condensed counterions are responsible for entropy-driven RNA collapse, analogous to the role of the hydrophobic effect in protein folding.  相似文献   

17.
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   

18.
Are lipid interactions with membrane proteins best described in terms of the physical properties of the lipid bilayer or in terms of direct molecular interactions between particular lipid molecules and particular sites on a protein? A molecular interpretation is more challenging because it requires detailed knowledge of the 3D structure of a membrane protein, but recent studies have suggested that a molecular interpretation is necessary. Here, the idea is explored that lipid molecules modify the ways that transmembrane α-helices pack into bundles, by penetrating between the helices and by binding into clefts between the helices, and that these effects on helix packing will modulate the activity of a membrane protein.  相似文献   

19.
Hia is a trimeric autotransporter found in the outer membrane of Haemphilus influenzae. The X-ray structure of Hia translocator domain revealed each monomer to consist of an α-helix connected via a loop to a 4-stranded β-sheet, thus the topology of the trimeric translocator domain is a 12-stranded β-barrel containing 3 α-helices that protrude from the mouth of the β-barrel into the extracellular medium. Molecular dynamics simulations of the Hia monomer and trimer have been employed to explore the interactions between the helices, β-barrel and connecting loops that may contribute to the stability of the trimer. In simulations of the Hia monomer we show that the central α-helix may stabilise the fold of the 4-stranded β-sheet. In simulations of the Hia trimer, a H-bond network involving residues in the β-barrel, α-helices and loops has been identified as providing stability for the trimeric arrangement of the monomers. Glutamine residues located in the loops connecting the α-helices to the β-barrel are orientated in a triangular arrangement such that each forms 2 hydrogen bonds to each of the corresponding glutamines in the other loops. In the absence of the loops, the β‐barrel becomes distorted. Simulations show that while the trimeric translocator domain β-barrel is inherently flexible, it is unlikely to accommodate the passenger domain in a folded conformation. Simulations of Hia in an asymmetric model of the outer membrane have revealed membrane–protein interactions that anchor the protein within its native membrane environment.  相似文献   

20.
The effects of three cationic surfactants, dodecylammonium chloride (DAC), dodecyldimethylammonium chloride (DDAC), and dodecyldimethylammonium chloride (DTDAC), on the conformation of poly(L -glutamic acid) and at neutral pH were examined by CD. The maximum extent of the α-helix induction occurs for each surfactant when the mixing ration is about unity. Different effects specific to each surfactant, as described below, appear in the range of mixing ratios larger than that required for the maximum induction. In the case of DTAC, the α-helices disintegrate into random coils. In the case of DDAC, the aggregation of α-helices takes place eventually leading to precipitation. Solubilization of the precipitates occurs at high mixing ratios. The most complex behavior is seen in the case of DAC; aggregation of α-helices occurs only to a small extent and the formation of a small complex predominates over aggregation takes place again as DAC concentration increases further. Induction of the α-helix is favored by dilution at a constant mixing ratio but is suppressed by the addition of NaCl.  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